id
stringlengths
27
33
source
stringclasses
1 value
format
stringclasses
1 value
text
stringlengths
13
1.81M
warning/0506/hep-th0506056.html
ar5iv
text
# Contents ## 1 Introduction and main results Classical supersymmetric solutions of supergravity theories have played a very important role in many advances in string theory for the past 15 years and are still the subject of much interest since they include, for example, backgrounds (possibly with branes and fluxes) for string model-building and supersymmetric objects such as black holes, supertubes and, as it has been discovered recently in Ref. , black rings. It is, thus, a very interesting problem to try to find or at least classify and characterize the supersymmetric solutions of (ideally all) supergravity theories. There have been many interesting results in the literature on this program, starting with the work of Gibbons and Hull in $`N=2,d=4`$ supergravity , completed in the seminal paper Ref. by Tod, who, starting from the Killing spinor equations (KSEs) of that theory and using all the integrability conditions and properties derived from them, assuming the existence on one Killing spinor, was able to find, for the first time, all the field configurations (metric and vector field strength) for which the KSEs could be solved. His classification included field configurations which may or may not satisfy the classical equations of motion. It was only 12 years later that a similar task was undertaken again by Tod, who in Ref. studied the supersymmetric solutions of pure, ungauged, $`N=4,d=4`$ supergravity, achieving a complete classification of the degenerate case (in which the Killing spinor gives rise to a null Killing vector) and only a partial classification of the non-degenerate case (in which the Killing spinor gives rise to a timelike Killing vector), since he had to assume a hypothesis of internal rigidity that he could not prove. The internally rigid cases were very interesting, though, since, as shown in Ref. they included all known the supersymmetric black-hole solutions of the theory, constructed by different methods and studied in Refs. -. By deformation of the supersymmetric black-hole solutions, the most general non-extremal black-hole solutions of the theory were constructed in Ref. . The program enjoyed a revival when a new maximally supersymmetric solution of $`N=2B,d=10`$ supergravity was discovered in Ref. , analogous maximally supersymmetric solutions of 11-dimensional and $`N=2,d=4`$ supergravity were rediscovered and additional maximally supersymmetric solutions of the same kind were found in 5 and 6 dimensions in Ref. . The classification of the maximally supersymmetric vacua of the 11- and 10-dimensional theories was completed in Refs. . It was then realized that we still had a very incomplete knowledge of the landscape of supersymmetric solutions of even the simplest supergravity theories and that new interesting supersymmetric solutions could be found by a systematic study of the solutions of the KSEs. This was done in Ref. for the minimal 5-dimensional supergravity, using a technique different from Tod’s, who used the Newman-Penrose formalism. In this work, the KSEs were translated into a set of differential equations on all the tensors that could be constructed as bilinears of the Killing spinors, which can be managed by more standard techniques. Several of the new solutions found in this work have had a great impact: a new maximally supersymmetric solution of Gödel type and the supersymmetric black rings <sup>3</sup><sup>3</sup>3More general black ring solutions have also been found in Refs. and generalizations that lead, for instance, to supersymmetric 4-dimensional rotating two- and one-black-hole solutions . This work was generalized to minimal gauged 5-dimensional supergravity in Ref. and then analogous results were obtained for minimal 6-dimensional supergravity in Refs. and for $`N=2,d=4`$ $`U(1)`$ gauged supergravity in Ref. . There is also extensive work on the 11-dimensional and $`N=2A,B,d=10`$ supergravities (see e.g. Refs. -), although a complete classification is still lacking. In this paper we return to the problem of finding all the supersymmetric configurations of $`N=4,d=4`$ supergravity, partially solved by Tod in Ref. . We use tensor methods, based on the bilinears of complex chiral spinors with $`SU(4)`$ indices, which allows us to keep manifest the S and T dualities of the theory at all stages in our analysis and in the field configurations, as it happens in the solutions studied in Ref. . The formalism used here can be used as starting point for the study of more complicated theories such as gauged and matter-coupled $`N=4,d=4`$ theories and there is work in progress in these directions. We are going to describe our main results in a moment but, before, it is worth explaining why $`N=4,d=4`$ supergravity is an interesting theory from the string theory point of view. The toroidal compactification of the heterotic string effective action ($`N=1,d=10`$ supergravity coupled to 16 vector multiplets) gives ungauged $`N=4,d=4`$ supergravity coupled to 22 (matter) vector multiplets and a consistent truncation of the matter vector multiplets gives the pure theory that we study here. Thus, all the solutions we will find are also solutions of the heterotic string effective action. The truncation preserves some of the $`SO(6,22;)`$ T duality symmetry and the theory is invariant under the continuous group $`SO(6)SU(4)`$ which naturally occurs as a hidden symmetry of the theory<sup>4</sup><sup>4</sup>4The first $`N=4,d=4`$ theory, constructed in Ref. had only $`SO(4)`$ invariance. We will work with the $`SU(4)`$ theory of Ref. . . The theory also has an S duality which manifest itself as a continuous $`SL(2,)`$ hidden symmetry. It was this symmetry which lead to the S duality conjectures in the corresponding superstring theory -. We will also keep this symmetry manifest at all stages in our analysis. Let us now describe our results for supersymmetric solutions, leaving the more general conditions for supersymmetric configurations which may or may not be solutions of the equations of motion. There are two types of supersymmetric solutions in $`N=4,d=4`$ supergravity admitting at least one Killing spinor $`ϵ_I`$, that can be characterized by the causal nature of the vector bilinear $`V^a=i\overline{ϵ}^I\gamma ^aϵ_I`$, which is always a non-spacelike Killing vector. Supersymmetric solutions are determined by a choice of 6 time-independent complex scalars $`M_{IJ}`$ and a complex scalar $`\tau `$ that in general may depend on the spatial coordinates $`x,z,z^{}`$. The $`M_{IJ}`$s have to satisfy two conditions: 1. Their matrix must have vanishing Pfaffian $$\epsilon ^{IJKL}M_{IJ}M_{KL}=0.$$ (1.1) 2. They must be such that the 1-form $`\xi `$ defined in Eq. (4.11) takes the form $$\xi =\pm \frac{i}{2}(_{\underset{¯}{z}}Udz_{\underset{¯}{z}^{}}Udz^{})+\frac{1}{2}d\lambda ,$$ (1.2) for some real functions $`U(z,z^{})`$ and $`\lambda (x,z,z^{})`$<sup>5</sup><sup>5</sup>5A general Ansatz that satisfies these two conditions is given in Eq. (4.64).. Observe that it is the function $`U`$ that makes $`\xi `$ non-trivial. $`\tau `$ and $`M_{IJ}`$ must satisfy the 3-dimensional differential equations $$_{\underset{¯}{i}}(e^{2i\lambda }A^{\underset{¯}{i}})e^{2i\lambda }[_{\underset{¯}{z}}(e^{2U})A_{\underset{¯}{z}^{}}_{\underset{¯}{z}^{}}(e^{2U})A_{\underset{¯}{z}}]=0,$$ (1.3) both for $$A=\frac{d\tau }{\mathrm{}\mathrm{m}\tau |M|^2},\mathrm{and}A=\frac{d[(\mathrm{}\mathrm{m}\tau )^{1/2}M^{IJ}]}{\mathrm{}\mathrm{m}\tau |M|^2},|M|^2=M^{IJ}M_{IJ},$$ (1.4) relative to the 3-dimensional metric $$\gamma _{\underset{¯}{i}\underset{¯}{j}}dx^{\underset{¯}{i}}dx^{\underset{¯}{j}}=dx^2+2e^{2U(z,z^{})}dzdz^{},$$ (1.5) whose triviality is associated to that of the connection $`\xi `$. Then, the metric is given by $$ds^2=|M|^2(dt+\omega )^2|M|^2(dx^2+2e^{2U}dzdz^{}),$$ (1.6) where $`\omega =\omega _{\underset{¯}{i}}dx^i`$ satisfies $$f_{ij}=4|M|^2ϵ_{ijk}\left(\xi _k\frac{_k\mathrm{}\mathrm{e}\tau }{4\mathrm{}\mathrm{m}\tau }\right),f_{\underset{¯}{i}\underset{¯}{j}}2_{[\underset{¯}{i}}\omega _{\underset{¯}{j}]}.$$ (1.7) again relative to the above 3-dimensional metric and the vector field strengths are given by $$F_{IJ}=\frac{1}{2|M|^2}\{\widehat{V}dE_{IJ}{}_{}{}^{}[\widehat{V}(\frac{\mathrm{}\mathrm{e}\tau }{\mathrm{}\mathrm{m}\tau }dE_{IJ}\frac{1}{\mathrm{}\mathrm{m}\tau }dB_{IJ})]\},$$ (1.8) where $$\begin{array}{ccc}\hfill \widehat{V}& =& \sqrt{2}|M|^2(dt+\omega ),\hfill \\ & & \\ \hfill E_{IJ}& =& 2\sqrt{2}(\mathrm{}\mathrm{m}\tau )^{1/2}(M_{IJ}+\stackrel{~}{M}_{IJ}),\hfill \\ & & \\ \hfill B_{IJ}& =& 2\sqrt{2}(\mathrm{}\mathrm{m}\tau )^{1/2}(\tau M_{IJ}+\tau ^{}\stackrel{~}{M}_{IJ}),\hfill \end{array}$$ (1.9) Examples of solutions corresponding to specific choices of $`M_{IJ}`$ and $`\tau `$ are given in Section 4.4, but it is clear that there are two different kinds of solutions which differ by the triviality of the connection $`\xi `$ and the 3-dimensional metric. The case in which $`\xi `$ is trivial was completely solved by Tod in Ref. . This case (called degenerate by Tod) was essentially solved by Tod in Ref. , but we study it here again for the sake of completeness and to refine his results. There are two subcases which we call $`A`$ and $`B`$ and which are associated to $`U(1)`$ holonomy in a null direction and in a pair of spacelike directions, respectively, and describe $`pp`$-waves and the stringy cosmic strings of Ref. . Each solution in this class is determined by 5 arbitrary functions of $`u`$: $`\varphi _I,\tau `$. Given these functions, the metric and vector field strengths are given by $$\begin{array}{ccc}\hfill ds^2& =& 2du[dv+K(u,z,z^{})du]2dzdz^{},\hfill \\ & & \\ \hfill F_{IJ}& =& \frac{1}{2}(_{IJ}+\frac{1}{2}\epsilon _{IJKL}^{KL})dudz^{},\hfill \end{array}$$ (1.10) where $$\begin{array}{ccc}\hfill _{IJ}& =& \frac{8\sqrt{2}}{(\mathrm{}\mathrm{m}\tau )^{1/2}}\dot{\varphi }_{[I}\varphi _{J]},\hfill \\ & & \\ \hfill 2_{\underset{¯}{z}}_{\underset{¯}{z}^{}}K& =& \frac{|\dot{\tau }|^2}{(\mathrm{}\mathrm{m}\tau )^2}+\frac{1}{16}\mathrm{}\mathrm{m}\tau ^2.\hfill \end{array}$$ (1.11) These are well-known solutions determined by a choice of (in this case) antiholomorphic function $`\tau =\tau (z^{})`$. The vector field strengths vanish<sup>6</sup><sup>6</sup>6These solutions are given in Ref. in different coordinates in which the metric functions have dependence on $`u`$, but this dependence can be eliminated. and the metric takes the form $$ds^2=2dudv2e^{2U}dzdz^{},e^{2U}=\mathrm{}\mathrm{m}(\tau ).$$ (1.12) As for the unbroken supersymmetries of these solutions, they all preserve generically $`1/4`$ of the supersymmetries. It is not easy to find generic conditions for the solutions to preserve $`1/2`$ (although this has been studied in special cases, see Ref. ). As for maximally supersymmetric solutions, we only expect Minkowski spacetime, since, otherwise, there would be another maximally supersymmetric solution of $`N=1,d=10`$ supergravity different from 10-dimensional Minkowski spacetime. The rest of this paper is devoted to proof these results. In Section 2 we describe in detail pure, ungauged, $`N=4,d=4`$ supergravity. In Section 3 we define the problem and equations that we want to solve and find the first consistency conditions. To go on, one has to consider separately the timelike and null cases. This is done in Sections 4 and 5, respectively. Our conventions are described in Appendix A and Appendix B contains all the algebraic identities satisfied by the products of tensors constructed as bilinears of chiral spinors, derived by Fierzing. ## 2 Pure, ungauged, $`N=4,d=4`$ supergravity The bosonic fields of $`N=4,d=4`$ supergravity multiplet are: 1. The Einstein metric $`g_{\mu \nu }`$. 2. The complex scalar $`\tau `$ that parametrizes an $`SL(2,)/U(1)`$ coset space. In terms of its real and imaginary parts (the axion $`a`$ and the dilaton $`\varphi `$) it is written $`\tau =a+ie^\varphi `$. 3. The 6 $`U(1)`$ vector fields whose complex combinations we label with an antisymmetric pair of $`SU(4)`$ indices $`A_{IJ\mu }`$, $`I,J=1,\mathrm{},4`$ and are subject to the reality constraint $$A_{IJ\mu }=\frac{1}{2}\epsilon _{IJKL}A^{KL}{}_{\mu }{}^{},$$ (2.1) where we rise and lower all $`SU(4)`$ indices by complex conjugation: $`A^{IJ}{}_{\mu }{}^{}(A_{IJ\mu })^{}`$. Their field strengths are $`F_{IJ}=dA_{IJ}`$ and are subject to the same reality constraint. The fermionic fields of this supermultiplet, which are always 4-component (complex) Weyl spinors, are 1. The 4 dilatini $`\chi _I`$, which, with lower $`SU(4)`$ indices, have positive chirality. 2. The 4 gravitini $`\psi _{I\mu }`$ which, with lower $`SU(4)`$ indices, have negative chirality. Complex conjugation raises the $`SU(4)`$ indices and reverses the chiralities. There are two global (hidden) symmetries in the ungauged theory: $`SU(4)SO(6)`$, associated to stringy T dualities and $`SL(2,)`$, which is associated to a stringy S duality - and leaves invariant the equations of motion but not the action. $`SU(4)`$ acts on all the fields in the obvious way: $$\chi ^I=U^I{}_{J}{}^{}\chi _{}^{J},\chi _I{}_{}{}^{}=\chi _J(U^{})^J{}_{I}{}^{},$$ (2.2) etc. The matrix $`\mathrm{\Lambda }=\left(\begin{array}{cc}a& b\\ c& d\end{array}\right)SL(2,)`$ acts on $`\tau `$ via fractional-linear transformations $$\tau ^{}=\frac{a\tau +b}{c\tau +d}.$$ (2.3) An alternative, linear, description of the action of $`\mathrm{\Lambda }SL(2,)`$ on $`\tau `$ can be made using the symmetric $`SL(2,)`$ matrix $$\frac{1}{\mathrm{}\mathrm{m}\tau }\left(\begin{array}{ccc}\hfill |\tau |^2& \mathrm{}\mathrm{e}\tau & \\ \hfill \mathrm{}\mathrm{e}\tau & 1& \end{array}\right).$$ (2.4) The fractional-linear transformations of $`\tau `$ are equivalent to the rule $$^{}=\mathrm{\Lambda }\mathrm{\Lambda }^T.$$ (2.5) Observe that the matrix $`Si\sigma ^2`$ is invariant under $`SL(2,)`$ transformations: $$\mathrm{\Lambda }S\mathrm{\Lambda }^T=S.$$ (2.6) The action of $`\mathrm{\Lambda }SL(2,)`$ on the vector fields is best described by defining the $`SL(2,)`$-dual $`\stackrel{~}{F}_{IJ}`$ of the field strength by $$\stackrel{~}{F}_{IJ}\tau F_{IJ}{}_{}{}^{+}+\tau ^{}F_{IJ}{}_{}{}^{}=\mathrm{}\mathrm{e}\tau F_{IJ}\mathrm{}\mathrm{m}\tau {}_{}{}^{}F_{IJ}^{}.$$ (2.7) Then, the pair $`\stackrel{~}{F}_{IJ},F_{IJ}`$ transforms as an $`SL(2,)`$ doublet, i.e. $$\stackrel{}{F}_{IJ}\left(\begin{array}{c}\stackrel{~}{F}_{IJ}\\ F_{IJ}\end{array}\right),\stackrel{}{F}_{IJ}^{}=\mathrm{\Lambda }\stackrel{}{F}_{IJ}.$$ (2.8) This implies for $`F_{IJ}^\pm `$ $$F_{IJ}^{}{}_{}{}^{+}=(c\tau +d)F_{IJ}{}_{}{}^{+},F_{IJ}^{}{}_{}{}^{}=(c\tau ^{}+d)F_{IJ}{}_{}{}^{}.$$ (2.9) Defining the phase of $`c\tau +d`$ by $$e^{2i\phi }\frac{c\tau +d}{c\tau ^{}+d},$$ (2.10) we find that, under $`SL(2,)`$ several fields and combinations of fields get a local $`U(1)`$ phase $$\begin{array}{cc}\chi _I^{}=e^{3i\phi /2}\chi _I,& \psi _{I\mu }^{}=e^{i\phi /2}\psi _{I\mu },\\ & \\ \left(\frac{_\mu \tau }{\mathrm{}\mathrm{m}\tau }\right)^{}=e^{2i\phi }\left(\frac{_\mu \tau }{\mathrm{}\mathrm{m}\tau }\right),& [(\mathrm{}\mathrm{m}\tau )^{1/2}F_{IJ}{}_{}{}^{\pm }{}_{\mu \nu }{}^{}]^{}=e^{\pm i\phi }[(\mathrm{}\mathrm{m}\tau )^{1/2}F_{IJ}{}_{}{}^{\pm }{}_{\mu \nu }{}^{}],\end{array}$$ (2.11) corresponding to $`U(1)`$ charges $`3,1,4`$ and $`\pm 2`$ respectively. The combination $$Q_\mu \frac{1}{4}\frac{_\mu \mathrm{}\mathrm{e}\tau }{\mathrm{}\mathrm{m}\tau },$$ (2.12) transforms as a $`U(1)`$ gauge field, $`Q_\mu ^{}=Q_\mu +\frac{1}{2}_\mu \phi `$ and this allows us to define a $`U(1)`$-covariant derivative $$𝒟_\mu =_\mu iqQ_\mu ,$$ (2.13) acting on fields with $`U(1)`$ charge $`q`$. Complex conjugation reverses chirality and these $`U(1)`$ charges. The action for the bosonic fields is $$S=d^4x\sqrt{|g|}\left[R+\frac{1}{2}\frac{_\mu \tau ^\mu \tau ^{}}{(\mathrm{}\mathrm{m}\tau )^2}\frac{1}{16}\mathrm{}\mathrm{m}\tau F^{IJ\mu \nu }F_{IJ\mu \nu }\frac{1}{16}\mathrm{}\mathrm{e}\tau F^{IJ\mu \nu }{}_{}{}^{}F_{IJ\mu \nu }^{}\right].$$ (2.14) It is useful to introduce the following notation for the equations of motion of the bosonic fields: $$_a{}_{}{}^{\mu }\frac{1}{2\sqrt{|g|}}\frac{\delta S}{\delta e^a_\mu },\frac{2\mathrm{}\mathrm{m}\tau }{\sqrt{|g|}}\frac{\delta S}{\delta \tau },^{IJ\mu }\frac{8}{\sqrt{|g|}}\frac{\delta S}{\delta A_{IJ\mu }}.$$ (2.15) Then, the equations of motion take the form $`_{\mu \nu }`$ $`=`$ $`G_{\mu \nu }+\frac{1}{2}(\mathrm{}\mathrm{m}\tau )^2[_{(\mu }\tau _{\nu )}\tau ^{}\frac{1}{2}g_{\mu \nu }_\rho \tau ^\rho \tau ^{}]\frac{1}{4}\mathrm{}\mathrm{m}\tau F_{IJ}{}_{\mu }{}^{+}{}_{}{}^{\rho }F_{}^{IJ}{}_{\nu \rho }{}^{},`$ (2.16) $``$ $`=`$ $`𝒟_\mu \left({\displaystyle \frac{^\mu \tau ^{}}{\mathrm{}\mathrm{m}\tau }}\right)\frac{i}{8}\mathrm{}\mathrm{m}\tau F^{IJ+\rho \sigma }F_{IJ}{}_{}{}^{+}{}_{\rho \sigma }{}^{},`$ (2.17) $`^{IJ\mu }`$ $`=`$ $`_\nu {}_{}{}^{}\stackrel{~}{F}_{}^{IJ\nu \mu }.`$ (2.18) The Maxwell equation $`^{IJ\mu }`$ transforms as an $`SL(2,)`$ doublet together with the Bianchi identity which we denote for convenience $`^{IJ\mu }`$ $$^{IJ\mu }_\nu {}_{}{}^{}F_{}^{IJ\nu \mu }.$$ (2.19) It is easy to see that the combinations $$\frac{_{IJ}{}_{}{}^{\mu }\tau ^{}_{IJ}^\mu }{(\mathrm{}\mathrm{m}\tau )^{1/2}},\frac{_{IJ}{}_{}{}^{\mu }\tau _{IJ}^\mu }{(\mathrm{}\mathrm{m}\tau )^{1/2}},$$ (2.20) have $`U(1)`$ charges $`+2`$ and $`2`$, respectively. The equation of motion of the complex scalar $``$ has $`U(1)`$ charge $`+4`$ and the Einstein equation is neutral. For vanishing fermions, the supersymmetry transformation rules of the gravitini and dilatini, generated by 4 spinors $`ϵ_I`$ of negative chirality and $`U(1)`$ charge $`+1`$, are $`\delta _ϵ\psi _{I\mu }`$ $`=`$ $`𝒟_\mu ϵ_I\frac{i}{2\sqrt{2}}(\mathrm{}\mathrm{m}\tau )^{1/2}F_{IJ}{}_{}{}^{+}{}_{\mu \nu }{}^{}\gamma _{}^{\nu }ϵ^J,`$ (2.21) $`\delta _ϵ\chi _I`$ $`=`$ $`\frac{1}{2\sqrt{2}}{\displaystyle \frac{\overline{)}\tau }{\mathrm{}\mathrm{m}\tau }}ϵ_I\frac{1}{8}(\mathrm{}\mathrm{m}\tau )^{1/2}\overline{)}F_{IJ}{}_{}{}^{}ϵ_{}^{J}.`$ (2.22) We also need the supersymmetry transformation rules of the bosonic bosonic fields, which take the form $`\delta _ϵe_\mu ^a`$ $`=`$ $`\frac{i}{4}(\overline{ϵ}^I\gamma ^a\psi _{I\mu }+\overline{ϵ}_I\gamma ^a\psi ^I{}_{\mu }{}^{}),`$ (2.23) $`\delta _ϵ\tau `$ $`=`$ $`\frac{i}{\sqrt{2}}\mathrm{}\mathrm{m}\tau \overline{ϵ}^I\chi _I,`$ (2.24) $`\delta _ϵA_{IJ\mu }`$ $`=`$ $`{\displaystyle \frac{\sqrt{2}}{(\mathrm{}\mathrm{m}\tau )^{1/2}}}[\overline{ϵ}_{[I}\psi _{J]\mu }+\frac{i}{\sqrt{2}}\overline{ϵ}_{[I}\gamma _\mu \chi _{J]}+\frac{1}{2}ϵ_{IJKL}\left(\overline{ϵ}^K\psi ^L{}_{\mu }{}^{}+\frac{i}{\sqrt{2}}\overline{ϵ}^K\gamma _\mu \chi ^L\right)].`$ (2.25) ## 3 Supersymmetric configurations: general setup Our goal is to find all the purely bosonic field configurations of $`N=4,d=4`$ supergravity $`\{g_{\mu \nu },A_{IJ\mu },\tau ,\psi _{I\mu }=0,\chi _I=0\}`$ which are supersymmetric, i.e. invariant under, at least, one supersymmetry transformation generated by a supersymmetry parameter $`ϵ_I(x)`$. Since the supersymmetry variations of the bosonic fields are odd in fermion fields, these transformations will always vanish, but the supersymmetry variations of the fermions, for vanishing fermions, Eqs. (2.21), may only vanish for special supersymmetry parameters $`ϵ_I(x)`$ (Killing spinors) that solve the Killing spinor equations (KSEs) $`\delta _ϵ\psi _{I\mu }=𝒟_\mu ϵ_I\frac{i}{2\sqrt{2}}(\mathrm{}\mathrm{m}\tau )^{1/2}F_{IJ}{}_{}{}^{+}{}_{\mu \nu }{}^{}\gamma _{}^{\nu }ϵ^J`$ $`=`$ $`0,`$ (3.1) $`2\sqrt{2}\delta _ϵ\chi _I={\displaystyle \frac{\overline{)}\tau }{\mathrm{}\mathrm{m}\tau }}ϵ_I\frac{1}{2\sqrt{2}}(\mathrm{}\mathrm{m}\tau )^{1/2}\overline{)}F_{IJ}{}_{}{}^{}ϵ_{}^{J}`$ $`=`$ $`0.`$ (3.2) For a known bosonic field configuration these are, respectively differential and algebraic equations for the Killing spinor, which may or may not exist. We want to find precisely for which bosonic field configurations these equations do have at least one solution $`ϵ_I`$. Our procedure will consist in assuming the existence of such a solution and derive consistency conditions for the field configurations. We shall be talking most of the time about supersymmetric field configurations. These may or may not be solutions of the classical equations of motion. There are several conceptual and practical advantages in doing so. First of all, we would like to emphasize the fact that supersymmetry does not imply by itself that the equations of motion are solved, although in general it considerably simplifies the task of solving them. Secondly, it is sometimes useful to consider that there are external sources for the fields, out of the regions in which we are solving the equations of motion. Including those regions with sources implies staying off-shell. Finally, the off-shell equations of motion of theories with gauge symmetries obey certain gauge identities. In theories with local supersymmetry and for field configurations admitting Killing spinors, the gauge identities are known as Killing spinor identities (KSIs) and can be used either to reduce the number of equations to be explicitly checked or, having at hands all the off-shell equations of motion of certain field configuration as we will, they can be used as a consistency check that it is a supersymmetric field configuration. Since these identities are the first consistency conditions that can be derived from the KSEs, we are going to derive them in the next section. We are also going to see that they are related to the integrability conditions of the KSEs. then, in Section 3.2 we are going to explain the strategy that we will follow to find all the supersymmetric configurations. ### 3.1 Killing Spinor Identities (KSIs) and integrability conditions of the Killing spinor equations Using the supersymmetry transformation rules of the bosonic fields Eqs. (2.23,2.24) and (2.25) we can derive relations between the (off-shell) equations of motion of the bosonic fields that are satisfied by any field configuration $`\{e^a{}_{\mu }{}^{},A_{IJ\mu },\tau \}`$ admitting Killing spinors . These KSIs take, for this theory, the form $`i\overline{ϵ}^I\gamma ^a_a{}_{}{}^{\mu }+{\displaystyle \frac{1}{\sqrt{2}(\mathrm{}\mathrm{m}\tau )^{1/2}}}\overline{ϵ}_J^{\mu JI}`$ $`=`$ $`0,`$ (3.3) $`\overline{ϵ}^I+{\displaystyle \frac{1}{\sqrt{2}(\mathrm{}\mathrm{m}\tau )^{1/2}}}\overline{ϵ}_J\overline{)}^{JI}`$ $`=`$ $`0.`$ (3.4) Observe that it is implicitly assumed that the Bianchi identities are identically satisfied, i.e. $$_{IJ}{}_{}{}^{\mu }=0,$$ (3.5) and, therefore, these identities are not $`SL(2,)`$-covariant. We may have to take this point into account when comparing with the equations that we will actually find, but we can also find (with considerably more effort) the $`SL(2,)`$-covariant relations between the equations of motion from the integrability conditions of the Killing spinor equations (3.1) and (3.2). Thus, acting with $`𝒟_\mu `$ on the Eq. (3.1) using both Eq. (3.1) and Eq. (3.2) and antisymmetrizing on the vector indices we get $$\begin{array}{ccc}\hfill 𝒟_{[\mu }\delta _ϵ\psi _{I\nu ]}& =& \frac{1}{8}\frac{_{[\mu }\tau _{\nu ]}\tau ^{}}{(\mathrm{}\mathrm{m}\tau )^2}ϵ_I\hfill \\ & & \\ & & \frac{1}{8}\{R_{\mu \nu }{}_{}{}^{ab}\delta _{I}^{}{}_{}{}^{K}\mathrm{}\mathrm{m}\tau F_{IJ}{}_{[\mu }{}^{+}{}_{}{}^{a}F_{}^{KJ}{}_{\nu ]}{}^{}{}_{}{}^{b}\}\gamma _{ab}ϵ_K\hfill \\ & & \\ & & +\frac{1}{4\sqrt{2}}(\mathrm{}\mathrm{m}\tau )^{1/2}\{F_{IJ}{}_{}{}^{+}{}_{\rho [\nu }{}^{}_{\mu ]}^{}\tau 2i\mathrm{}\mathrm{m}\tau _{[\mu |}F_{IJ}{}_{}{}^{+}{}_{|\nu ]}{}^{}\}\gamma ^\rho ϵ^J\hfill \\ & & \\ & =& 0.\hfill \end{array}$$ (3.6) To extract from this integrability condition a relation between the equations of motion we act with $`\gamma ^\nu `$ from the left. We get $$4\gamma ^\nu 𝒟_{[\mu }\delta _ϵ\psi _{I\nu ]}=(_{\mu \nu }\frac{1}{2}g_{\mu \nu }_\sigma {}_{}{}^{\sigma })\gamma ^\nu ϵ_I\frac{i}{2\sqrt{2}(\mathrm{}\mathrm{m}\tau )^{1/2}}(\overline{)}_{IJ}\tau ^{}\overline{)}_{IJ})\gamma _\mu ϵ^J=0.$$ (3.7) Acting now with $`\gamma ^\mu `$ and using the result to eliminate $`_\sigma ^\sigma `$ we get, finally the $`SL(2,)`$-covariantization of the KSIs Eq. (3.3) $$^\mu {}_{a}{}^{}\gamma _{}^{a}ϵ_I\frac{i}{\sqrt{2}(\mathrm{}\mathrm{m}\tau )^{1/2}}(_{IJ}{}_{}{}^{\mu }\tau ^{}_{IJ}{}_{}{}^{\mu })ϵ^J=0.$$ (3.8) Similarly, the $`SL(2,)`$-covariantization of the KSIs Eq. (3.3) can be obtained by calculating $`2\sqrt{2}\overline{)}𝒟\delta _ϵ\chi _I=0`$ and takes the form $$^{}ϵ_I\frac{1}{\sqrt{2}(\mathrm{}\mathrm{m}\tau )^{1/2}}(\overline{)}_{IJ}\tau \overline{)}_{IJ})ϵ^J=0.$$ (3.9) These two identities are now manifestly $`SL(2,)`$-covariant<sup>7</sup><sup>7</sup>7See the paragraph after Eq. (2.20).. The comparison with our results will be easier if we multiply these equations by gamma matrices and conjugate spinors $`\overline{ϵ}_K`$ and $`\overline{ϵ}^K`$ from the left, to derive relations involving spinor bilinears. In the case in which the vector $`V^a`$ is timelike, we get $`^{ab}\frac{1}{2}\mathrm{}\mathrm{m}V^aV^b{\displaystyle \frac{1}{\sqrt{2}}}(\mathrm{}\mathrm{m}\tau )^{1/2}\mathrm{}\mathrm{m}(M^{IJ}_{IJ}{}_{}{}^{a})V^b`$ $`=`$ $`0,`$ (3.10) $`^{}V^a{\displaystyle \frac{i}{\sqrt{2}(\mathrm{}\mathrm{m}\tau )^{1/2}}}M^{IJ}(_{IJ}{}_{}{}^{a}\tau _{IJ}{}_{}{}^{a})`$ $`=`$ $`0,`$ (3.11) $`\mathrm{}\mathrm{m}[M_{IJ}(_{IJ}{}_{}{}^{a}\tau ^{}_{IJ}{}_{}{}^{a})]`$ $`=`$ $`0.`$ (3.12) Observe that the first equation implies the off-shell vanishing of all the Einstein equations with one or two spacelike components. Further, the Einstein equation is automatically satisfied when the Maxwell, Bianchi and complex scalar equations are satisfied. When $`V^a`$ is null (we denote it by $`l^a`$), all the spinors $`ϵ_I`$ are proportional and we can use the parametrization of Eq. (B.26) in Eqs. (3.8) and (3.9). Contracting with $`\varphi ^I`$ using the normalization Eq. (B.27) and with the conjugate spinors $`\overline{ϵ},\overline{ϵ^{}},\overline{\eta },\overline{\eta ^{}}`$, where $`\eta `$ is an auxiliary spinor with normalization Eq. (B.33), we arrive at the identities $`(^\mu {}_{a}{}^{}\frac{1}{2}e_a{}_{}{}^{\mu }_{}^{\rho }{}_{\rho }{}^{})l^a=(^\mu {}_{a}{}^{}\frac{1}{2}e_a{}_{}{}^{\mu }_{}^{\rho }{}_{\rho }{}^{})m^a`$ $`=`$ $`0,`$ (3.13) $``$ $`=`$ $`0,`$ (3.14) $`(_{IJ}{}_{}{}^{\mu }\tau ^{}_{IJ}{}_{}{}^{\mu })\varphi ^J`$ $`=`$ $`0.`$ (3.15) where the null complex vectors are defined in Eq. (B.34). Observe that in this case supersymmetry implies that the scalar equations of motion must be automatically satisfied. ### 3.2 Solving the Killing spinor equations The procedure we will follow to find the field configurations for which the KSEs admit at least one solution will be the following: 1. In Section 3.3 we are going to reexpress the KSEs as differential and algebraic equations for the bilinears (scalars, vectors and 2-forms, see Appendix B) built with the Killing spinors. Solving the equations for all the bilinears is essentially equivalent to solving the KSEs. 2. In Section 3.4 we are going to find, among the bilinears, a Killing vector $`V^\mu `$ and decompose the vector field strengths w.r.t. to it computing $`V^\rho F_{IJ}{}_{}{}^{+}_{\mu \rho }`$ or $`V^\rho F_{IJ}{}_{}{}^{}_{\mu \rho }`$ in terms of the scalar bilinears and $`\tau `$ and then using, Eqs. (A.16) if $`V`$ is timelike and Eqs. (A.19) if $`V`$ is null. These two cases have to be studied separately. One of the reasons is that, in the null case, the field strength is not completely determined by its contractions with $`V`$, but there are more differences that we are going to explain shortly and require a completely separate analysis. 3. In the timelike case, studied in Section 4 we will 1. Substitute the expressions of the field strengths in the algebraic KSEs ($`\delta _ϵ\chi _I=0`$) to check that it is completely solved. 2. Substitute into the equations of motion and we will check whether the KSIs Eqs. (3.10,3.11,3.12) are indeed satisfied or there are additional conditions to be imposed. This is done in two steps: first we substitute into the equations of motion of the vector fields and the complex scalar which we have already expressed in terms of the bilinears in Section 4.1 and then, after we specify the form of the metric in terms of the bilinears, we substitute into the Einstein equations in Section 4.2. Then we check the KSIs. 3. Substitute, finally, into the differential KSEs ($`\delta _ϵ\psi _{I\mu }=0`$) to solve it finding additional conditions on the bilinears and the form of the Killing spinors in Section 4.3. The timelike case will be completely solved by then and we will study some examples. In the null case, which was completely solved by Tod, 1. As explained in Appendix B all the spinors $`ϵ_I`$ are proportional $`ϵ_I=\varphi _Iϵ`$ and we use first this information in the KSEs to obtain separate equations for the coefficients $`\varphi _I`$ and the spinor $`ϵ`$. This requires the introduction of a $`U(1)`$ connection $`\zeta `$ that covariantizes the equations with respect to (opposite) local changes of phase of $`\varphi _I`$ and $`ϵ`$. 2. All the vectors bilinears are also proportional to the Killing vector $`V^a`$ which we rename here $`l^a`$. It is convenient to introduce an auxiliary spinor to build independent vector bilinears that constitute a null tetrad. The KSEs only give partial information about the derivatives of these vectors, except for $`l^a`$, which is built with $`ϵ`$ and is always covariantly constant, the very definition of a $`pp`$-wave space . 3. Although the vector field strengths and the derivatives of the vector bilinears are not completely determined, it is possible to extract information constructing the equations of motion and imposing the KSI. In particular we find that the $`U(1)`$ connection $`\zeta `$ is trivial. 4. There are two different cases to be considered ($`A`$ and $`B`$) which are essentially solved by solving first the integrability constraints. ### 3.3 Killing equations for the bilinears We start with the equations $`\delta _ϵ\chi _I=0`$. We just have to multiply the from the right with gamma matrices and Dirac conjugates of Killing spinors. We have, in particular, from $`\overline{ϵ}^K\delta _ϵ\chi _I=0`$ $$V^K{}_{I}{}^{}\tau \frac{i}{2\sqrt{2}}(\mathrm{}\mathrm{m}\tau )^{3/2}F_{IJ}{}_{}{}^{}\mathrm{\Phi }^{KJ}=0,$$ (3.16) and, from $`\overline{ϵ}^K\gamma _\rho \delta _ϵ\chi _I=0`$ $$F_{IJ}{}_{}{}^{}{}_{\rho \sigma }{}^{}V_{}^{J}{}_{K}{}^{}{}_{}{}^{\sigma }+\frac{i}{\sqrt{2}}(\mathrm{}\mathrm{m}\tau )^{3/2}(M_{IK}_\rho \tau \mathrm{\Phi }_{IK\rho }{}_{}{}^{\mu }_{\mu }^{}\tau )=0.$$ (3.17) It is possible to derive more Killing equations for the bilinears from the dilatini supersymmetry rule, but it will not be necessary. Let us turn to the gravitini supersymmetry rules. Now we apply $`SL(2,)`$-covariant derivative on the bilinears and use $`\delta _ϵ\psi _{I\mu }=0`$ to reexpress $`𝒟_\mu ϵ_I`$. We get $`𝒟_\mu M_{IJ}`$ $`=`$ $`\frac{1}{\sqrt{2}}(\mathrm{}\mathrm{m}\tau )^{1/2}F_{K[I|}{}_{}{}^{+}{}_{\mu \nu }{}^{}V_{}^{K}{}_{|J]}{}^{}{}_{}{}^{\nu },`$ (3.18) $`𝒟_\mu V^I_{J\nu }`$ $`=`$ $`\frac{1}{2\sqrt{2}}(\mathrm{}\mathrm{m}\tau )^{1/2}[M_{KJ}F^{KI}{}_{\mu \nu }{}^{}+M^{IK}F_{JK}{}_{}{}^{+}_{\mu \nu }`$ (3.19) $`\mathrm{\Phi }_{KJ(\mu }{}_{}{}^{\rho }F_{}^{KI}{}_{\nu )\rho }{}^{}\mathrm{\Phi }^{IK}{}_{(\mu |}{}^{}{}_{}{}^{\rho }F_{KI}^{}{}_{}{}^{+}{}_{|\nu )\rho }{}^{}],`$ $`𝒟_\mu \mathrm{\Phi }_{IJ\mu \nu }`$ $`=`$ $`\frac{1}{2\sqrt{2}}(\mathrm{}\mathrm{m}\tau )^{1/2}[2g_{\mu [\nu |}F_{KI}{}_{}{}^{+}{}_{|\rho ]\alpha }{}^{}V_{}^{K}{}_{J}{}^{}{}_{}{}^{\alpha }+2F_{KI}{}_{}{}^{+}{}_{\nu \rho }{}^{}V_{}^{K}_{J\mu }`$ (3.20) $`3F_{KI}{}_{}{}^{+}{}_{[\mu \nu |}{}^{}V_{}^{K}{}_{J|\rho ]}{}^{}+(IJ)].`$ ### 3.4 First consequences and general results Contracting the free indices in Eqs. (3.19) and (3.16) it is immediate to see that $`V^\mu V^I{}_{I}{}^{}^\mu `$ is a (non-spacelike, Eq. (B.15)) Killing vector and $$V^\mu _\mu \tau =0.$$ (3.21) It is also immediate to prove that $$_\mu V^I{}_{J}{}^{}{}_{}{}^{\mu }=0.$$ (3.22) Let us now consider the implications of the reality constraint of the vector field strengths on the contraction $`F_{KI}{}_{}{}^{+}{}_{\mu \nu }{}^{}V_{}^{K}{}_{J}{}^{}^\nu `$: $$F_{KI}{}_{}{}^{+}{}_{\mu \nu }{}^{}V_{}^{K}{}_{J}{}^{}{}_{}{}^{\nu }=\frac{1}{2}\epsilon _{KIML}(F_{ML}{}_{}{}^{}{}_{\mu \nu }{}^{})^{}V^K{}_{J}{}^{}{}_{}{}^{\nu }.$$ (3.23) Taking the $`SU(4)`$ dual in both sides of this equation and taking into account the reality properties of the vectors $`V^K{}_{J}{}^{}^\nu `$, we get $$\frac{1}{2}\epsilon ^{SRIJ}F_{KI}{}_{}{}^{+}{}_{\mu \nu }{}^{}V_{}^{K}{}_{J}{}^{}{}_{}{}^{\nu }=\frac{1}{2}[F_{SR}{}_{}{}^{}{}_{\mu \nu }{}^{}V_{}^{\nu }+2F_{J[S|}{}_{}{}^{}{}_{\mu \nu }{}^{}V_{}^{J}{}_{|R]}{}^{}{}_{}{}^{\nu }]^{},$$ (3.24) from which we get $$F_{SR}{}_{}{}^{}{}_{\mu \nu }{}^{}V_{}^{\nu }=2F_{J[S|}{}_{}{}^{}{}_{\mu \nu }{}^{}V_{}^{J}{}_{|R]}{}^{}{}_{}{}^{\nu }[\epsilon ^{SRIJ}F_{KI}{}_{}{}^{+}{}_{\mu \nu }{}^{}V_{}^{K}{}_{J}{}^{}{}_{}{}^{\nu }]^{}.$$ (3.25) The first and second terms in the r.h.s. of this equation can be rewritten in terms of scalars using the antisymmetric part of Eq. (3.17) and the complex conjugate of Eq. (3.18). We get, at last, $$F_{SR}{}_{}{}^{}{}_{\mu \nu }{}^{}V_{}^{\nu }=\frac{\sqrt{2}i}{(\mathrm{}\mathrm{m}\tau )^{3/2}}M_{SR}_\mu \tau \frac{\sqrt{2}}{(\mathrm{}\mathrm{m}\tau )^{1/2}}\epsilon _{SRIJ}𝒟_\mu M^{IJ}.$$ (3.26) The complex conjugate of this equation gives us $`F^{SR+}{}_{\mu \nu }{}^{}V_{}^{\nu }`$ and, taking the $`SU(4)`$-dual we get $`F_{IJ}{}_{}{}^{+}{}_{\mu \nu }{}^{}V_{}^{\nu }`$ etc. From this equation, contracting the free index with $`V^\mu `$ and using Eq. (3.21) we get immediately $$V^\mu _\mu M_{IJ}=0.$$ (3.27) Now, the use that we make of this result and the subsequent analysis will depend on the causal nature if the non-spacelike vector $`V^\mu `$. We must distinguish between two cases: the case in which it is timelike, which we consider in section 4 and the case in which it is null (and we rename it $`l^\mu `$), which we consider in section 5. ## 4 The timelike case If $`V^2=2M^{IJ}M_{IJ}2|M|^20`$ we can use Eq. (3.26) to express $`F_{IJ}^{}`$ entirely in terms of scalars, their derivatives, and $`V_\mu `$ using Eq. (A.16): $$F_{SR}{}_{}{}^{}=\frac{1}{\sqrt{2}|M|^2(\mathrm{}\mathrm{m}\tau )^{1/2}}\{[i\frac{M_{SR}}{(\mathrm{}\mathrm{m}\tau )}d\tau +\epsilon _{SRIJ}𝒟M^{IJ}]\widehat{V}i{}_{}{}^{}\left[\mathrm{}\right]\}.$$ (4.1) Here we have added a hat to $`V`$ to denote the differential form $`\widehat{V}V_\mu dx^\mu `$ and distinguish its norm. It can be seen that this form of $`F_{SR}^{}`$ satisfies identically all the Killing spinor equations $`\delta _ϵ\chi _I=0`$, that we can consider solved. To solve the equations of motion it is convenient to have directly $`F_{IJ}`$ and its $`SL(2,)`$-dual $`\stackrel{~}{F}_{IJ}`$. Their expressions are, actually, somewhat simpler due to the following property: if $`dF=0`$ (which is the equation satisfied by $`F_{IJ}`$ and $`\stackrel{~}{F}_{IJ}`$) and $`\mathrm{\pounds }_VF=0`$ then $`_{[\mu }(F_{\nu ]\rho }V^\rho )=0`$ and, locally, $`F_{\nu \rho }V^\rho =_\nu E`$ for some scalar potential $`E`$. Thus, following Tod , we define $$_\mu E_{IJ}V^\nu F_{IJ\nu \mu },_\mu B_{IJ}V^\nu \stackrel{~}{F}_{IJ\nu \mu },$$ (4.2) and, using the above form of $`F_{IJ}^{}`$ Eq. (4.1) we find $$\begin{array}{ccc}\hfill E_{IJ}& =& 2\sqrt{2}(\mathrm{}\mathrm{m}\tau )^{1/2}(M_{IJ}+\stackrel{~}{M}_{IJ}),\hfill \\ & & \\ \hfill B_{IJ}& =& 2\sqrt{2}(\mathrm{}\mathrm{m}\tau )^{1/2}(\tau M_{IJ}+\tau ^{}\stackrel{~}{M}_{IJ}),\hfill \end{array}$$ (4.3) where $`\stackrel{~}{F}_{IJ}`$ $`=`$ $`V^2\{\widehat{V}dB_{IJ}+{}_{}{}^{}[\widehat{V}({\displaystyle \frac{\mathrm{}\mathrm{e}\tau }{\mathrm{}\mathrm{m}\tau }}dB_{IJ}{\displaystyle \frac{|\tau |^2}{\mathrm{}\mathrm{m}\tau }}dE_{IJ})]\},`$ (4.4) $`F_{IJ}`$ $`=`$ $`V^2\{\widehat{V}dE_{IJ}{}_{}{}^{}[\widehat{V}({\displaystyle \frac{\mathrm{}\mathrm{e}\tau }{\mathrm{}\mathrm{m}\tau }}dE_{IJ}{\displaystyle \frac{1}{\mathrm{}\mathrm{m}\tau }}dB_{IJ})]\}.`$ (4.5) It is worth spending a moment in checking the consistency of these results. By definition, $`B_{IJ}`$ and $`E_{IJ}`$ must transform under $`SL(2,)`$ as $`\stackrel{~}{F}_{IJ}`$ and $`F_{IJ}`$, i.e. as a doublet: $$\stackrel{}{E}_{IJ}\left(\begin{array}{c}B_{IJ}\\ E_{IJ}\end{array}\right),\stackrel{}{E}_{IJ}^{}=\mathrm{\Lambda }\stackrel{}{E}_{IJ}.$$ (4.6) We can check that this is consistent with Eqs. (4.4) and (4.5) by rewriting the last two equations in the manifestly $`SL(2,)`$-covariant form $$\stackrel{}{F}_{IJ}=V^2\{\widehat{V}d\stackrel{}{E}_{IJ}{}_{}{}^{}[\widehat{V}(Sd\stackrel{}{E}_{IJ})]\},$$ (4.7) on account of Eqs. (2.4,2.5) and (2.6). On the other hand, it is easy to check that the fact that $`\stackrel{}{E}_{IJ}`$ transforms as a doublet is consistent with the transformations rules of $`\tau `$ and $`M_{IJ}`$ alone and Eqs. (4.3). ### 4.1 Vector and scalar equations of motion Our next step consists in finding equations for $`M_{IJ}`$ and $`\tau `$ from the equations of motion using the decompositions of $`\stackrel{~}{F}_{IJ}`$ and $`F_{IJ}`$ Eqs. (4.4) and (4.5) in which these fields are written entirely in terms of those scalars and the Killing vector (1-form) $`V`$. In this process we are going to find derivatives of $`V`$, and we need to express these in terms of the scalars and $`V`$ itself. From Eq. (3.19) we find that $`V`$ satisfies the equation $$d\widehat{V}=\frac{1}{\sqrt{2}}(\mathrm{}\mathrm{m}\tau )^{1/2}[M^{IJ}F_{IJ}{}_{}{}^{+}+M_{IJ}F^{IJ}].$$ (4.8) Since $$M^{IJ}F_{IJ}{}_{}{}^{+}=\frac{\sqrt{2}M^{IJ}}{(\mathrm{}\mathrm{m}\tau )^{1/2}|M|^2}[𝒟M_{IJ}\widehat{V}+i{}_{}{}^{}(𝒟M_{IJ}\widehat{V})],$$ (4.9) we get $$d\widehat{V}=\frac{1}{|M|^2}\{d|M|^2\widehat{V}+i{}_{}{}^{}[(M^{IJ}𝒟M_{IJ}M_{IJ}𝒟M^{IJ})\widehat{V}]\}.$$ (4.10) It is also convenient to define the 1-form $`\xi `$ and the 2-form $`\mathrm{\Omega }`$ $`\xi `$ $``$ $`\frac{i}{4}|M|^2(M_{IJ}dM^{IJ}M^{IJ}dM_{IJ}),`$ (4.11) $`\mathrm{\Omega }`$ $``$ $`2|M|^2{}_{}{}^{}[(Q\xi )\widehat{V}].`$ (4.12) $`\xi `$ transforms under $`SL(2,)`$ as $$\xi ^{}=\xi +\frac{1}{2}d\phi ,$$ (4.13) i.e. as the $`U(1)`$ connection $`Q`$, which makes $`\mathrm{\Omega }`$ invariant. The connection $`\xi `$ is also orthogonal to $`V`$ and invariant under local rescalings of the scalar matrix $`M_{IJ}`$: $$\xi (\mathrm{\Lambda }(x)M_{IJ})=\xi (M_{IJ}),$$ (4.14) a property that we will exploit later on. Further, using Eq. (B.24) we can write the curvature of this connection in the form $$d\xi =\frac{i}{2}d\frac{M^{IJ}}{|M|}d\frac{M_{KL}}{|M|}[\delta _{IJ}{}_{}{}^{KL}𝒥^K{}_{[I}{}^{}𝒥_{}^{L}{}_{J]}{}^{}],$$ (4.15) that relates the triviality of $`\xi `$ with the constancy of the projection $`𝒥^I_J`$. Finally, it is convenient to rewrite the equations of motion of the vector and scalar fields in differential-form language<sup>8</sup><sup>8</sup>8We add hats to denote differential forms.: $`\widehat{\stackrel{}{}}^{IJ}`$ $``$ $`\stackrel{}{}^{IJ}{}_{\mu }{}^{}dx^\mu ={}_{}{}^{}d\stackrel{}{F}^{IJ}=\left(\begin{array}{c}\widehat{}^{IJ}\\ \widehat{}^{IJ}\end{array}\right),`$ (4.18) $`\widehat{}`$ $``$ $`\widehat{V},`$ (4.19) where $`\stackrel{}{}^{IJ}_\mu `$ is the $`SL(2,)`$ doublet formed by the Maxwell and Bianchi identities: $$\stackrel{}{}^{IJ\mu }\left(\begin{array}{c}^{IJ\mu }\\ ^{IJ\mu }\end{array}\right)=\left(\begin{array}{c}_\nu {}_{}{}^{}\stackrel{~}{F}_{}^{IJ\nu \mu }\\ _\nu {}_{}{}^{}F_{}^{IJ\nu \mu }\end{array}\right).$$ (4.20) Using the expressions that we have found for the Maxwell fields and their $`SL(2,)`$ duals and using the above equation for $`dV`$ rewritten in the form $$d\widehat{V}=\frac{d|M|^2}{|M|^2}\widehat{V}+2|M|^2\mathrm{\Omega },$$ (4.21) we find the following two equations for $`M_{IJ}`$ and $`\tau `$: $`{}_{}{}^{}\widehat{\stackrel{}{}}^{IJ}`$ $`=`$ $`\frac{1}{2}d{}_{}{}^{}[{\displaystyle \frac{Sd\stackrel{}{E}_{IJ}}{|M|^2}}\widehat{V}]+d\stackrel{}{E}_{IJ}\mathrm{\Omega },`$ (4.22) $`{\displaystyle \frac{{}_{}{}^{}\widehat{}_{}^{}}{|M|^2}}`$ $`=`$ $`𝒟{}_{}{}^{}[{\displaystyle \frac{d\tau }{|M|^2\mathrm{}\mathrm{m}\tau }}\widehat{V}]+2i{\displaystyle \frac{d\tau }{\mathrm{}\mathrm{m}\tau }}\mathrm{\Omega }+2i{\displaystyle \frac{\stackrel{~}{M}_{IJ}}{|M|^2}}d{}_{}{}^{}({\displaystyle \frac{dM^{IJ}}{|M|^2}}\widehat{V}).`$ (4.23) These equations can be now be combined (this is the reason behind the introduction of $`V`$ into the equation for $`\tau `$ and the use of differential forms) and simplified. Using the new variables $`N_{IJ}`$ defined by $$N_{IJ}=\sqrt{\mathrm{}\mathrm{m}\tau }M_{IJ},|N|^2=N^{IJ}N_{IJ}=\mathrm{}\mathrm{m}\tau |M|^2,$$ (4.24) we construct a new combination of equations that we call $`\widehat{a}^{IJ}`$ $$\widehat{a}^{IJ}\frac{1}{2\sqrt{2}\mathrm{}\mathrm{m}\tau }(\tau \widehat{}^{IJ}\widehat{}^{IJ})\frac{i}{2}\frac{(N^{IJ}+\stackrel{~}{N}^{IJ})}{|N|^2}\widehat{}^{},$$ (4.25) and, which, after some massaging, is going to have a much simpler form. To present in compact form the equations of motion we define these two equations $`n^{IJ}`$ $``$ $`(_\mu +4i\xi _\mu )\left({\displaystyle \frac{^\mu N^{IJ}}{|N|^2}}\right),`$ (4.26) $`e^{}`$ $``$ $`(_\mu +4i\xi _\mu )\left({\displaystyle \frac{^\mu \tau }{|N|^2}}\right),`$ (4.27) and, in terms of them, we have, switching again from differential form notation to tensor notation, $`a^{IJ}`$ $`=`$ $`n^{IJ}{\displaystyle \frac{N^{IJ}+\stackrel{~}{N}^{IJ}}{|N|^2}}\stackrel{~}{N}_{KL}n^{KL},`$ (4.28) $`^{IJa}`$ $`=`$ $`\sqrt{2}V^a\left\{{\displaystyle \frac{N^{IJ}+\stackrel{~}{N}^{IJ}}{|N|^2}}\mathrm{}\mathrm{e}i(a^{IJ}\stackrel{~}{a}^{IJ})\right\},`$ (4.29) $`^{IJa}`$ $`=`$ $`\sqrt{2}V^a\left\{{\displaystyle \frac{N^{IJ}+\stackrel{~}{N}^{IJ}}{|N|^2}}\mathrm{}\mathrm{e}(\tau )i(\tau ^{}a^{IJ}\tau \stackrel{~}{a}^{IJ})\right\}.`$ (4.30) $``$ $`=`$ $`|M|^2e+2i\stackrel{~}{N}^{KL}n_{KL}.`$ (4.31) The combination $`|N|^2d\tau `$ has $`U(1)`$ charge $`4`$ and, thus, the second equation is just a $`U(1)`$-covariant divergence, the covariant derivative being constructed with the $`\xi `$ connection. The first equation has a similar form and, although $`\frac{dN_{IJ}}{|N|^2}`$ does not transform covariantly under $`SL(2,)`$, the equation is $`SL(2,)`$-covariant up to terms proportional to the second equation. ### 4.2 Metric equations of motion These are equations for the scalars $`M_{IJ}`$ and $`\tau `$ and involve implicitly the spacetime metric, which is the only field not determined by them. We need to study now the Einstein equations and, to do it, it is convenient to choose coordinates adapted to the timelike Killing vector $`V`$. We define a time coordinate by $$V^\mu _\mu \sqrt{2}_t,$$ (4.32) and the metric takes the “conformastationary” form $$ds^2=|M|^2(dt+\omega )^2|M|^2\gamma _{\underset{¯}{i}\underset{¯}{j}}dx^idx^j,i,j=1,2,3,$$ (4.33) where $`\omega =\omega _{\underset{¯}{i}}dx^i`$ is a time-independent 1-form and $`\gamma _{\underset{¯}{i}\underset{¯}{j}}`$ is a time-independent (positive-definite!) metric on constant $`t`$ hypersurfaces<sup>9</sup><sup>9</sup>9The components of the connection and curvature of this metric can be found in Appendix C.. Since $`|M|`$ is in principle determined by the above equations, we only need to find equations for $`\omega `$ and $`\gamma `$. As usual, the equation for $`\omega `$ can be derived by comparing Eq. (4.21) for the 1-form $`\widehat{V}`$, with the exterior derivative of the expression for $`\widehat{V}`$ in the coordinates chosen $$\widehat{V}=\sqrt{2}|M|^2(dt+\omega ).$$ (4.34) The result is the equation $$d\omega =\frac{1}{\sqrt{2}}\mathrm{\Omega }=\frac{i}{2\sqrt{2}}|M|^4{}_{}{}^{}[(M^{IJ}𝒟M_{IJ}M_{IJ}𝒟M^{IJ})\widehat{V}].$$ (4.35) Using the conformastationary metric we can reduce all the equations to equations in the 3 spatial dimensions with the metric $`\gamma `$. To start with, the equations $`n^{IJ}`$ and $`e`$ defined in Eqs. (4.26) and (4.27) can be expressed in terms of $`n_{(3)}^{IJ}`$ $``$ $`(_{\underset{¯}{i}}+4i\xi _{\underset{¯}{i}})\left({\displaystyle \frac{^{\underset{¯}{i}}N^{IJ}}{|N|^2}}\right),`$ (4.36) $`e_{(3)}^{}`$ $``$ $`(_{\underset{¯}{i}}+4i\xi _{\underset{¯}{i}})\left({\displaystyle \frac{^{\underset{¯}{i}}\tau }{|N|^2}}\right),`$ (4.37) where all the objects are now 3-dimensional with metric $`\gamma `$, by $$n^{IJ}=|M|^2n_{(3)}^{IJ},e=|M|^2e_{(3)}.$$ (4.38) The equation (4.35) for the 1-form $`\omega `$ that enters the conformastationary metric reduces to $$f_{ij}=4|M|^2ϵ_{ijk}(\xi _kQ_k),f_{\underset{¯}{i}\underset{¯}{j}}2_{[\underset{¯}{i}}\omega _{\underset{¯}{j}]}.$$ (4.39) Then, we can express all the equations of motion in terms of these two equations plus the equation<sup>10</sup><sup>10</sup>10This equation should be compared with Eq. (4.15) in which the antisymmetric part of the same combination appears. $$e_{ij}R_{ij}(\gamma )2_{(i}\left(\frac{N^{IJ}}{|N|}\right)_{j)}\left(\frac{N_{KL}}{|N|}\right)(\delta ^{KL}{}_{IJ}{}^{}𝒥^K{}_{I}{}^{}𝒥_{}^{L}{}_{J}{}^{}),$$ (4.40) as follows: $`_{00}`$ $`=`$ $`|M|^2[|M|^2\mathrm{}\mathrm{m}e_{(3)}^{}2\mathrm{}\mathrm{e}(N_{KL}n_{(3)}^{KL})+\frac{1}{2}e_k{}_{}{}^{k}],`$ (4.41) $`_{0i}`$ $`=`$ $`0,`$ (4.42) $`_{ij}`$ $`=`$ $`|M|^2(e_{ij}\frac{1}{2}\delta _{ij}e_k{}_{}{}^{k}),`$ (4.43) $`^{IJa}`$ $`=`$ $`\sqrt{2}|M|^2V^a\left\{{\displaystyle \frac{N^{IJ}+\stackrel{~}{N}^{IJ}}{\mathrm{}\mathrm{m}\tau }}\mathrm{}\mathrm{e}e_{(3)}i(n_{(3)}^{IJ}\stackrel{~}{n}_{(3)}^{IJ})\right\},`$ (4.44) $`^{IJa}`$ $`=`$ $`\sqrt{2}|M|^2V^a\left\{{\displaystyle \frac{N^{IJ}+\stackrel{~}{N}^{IJ}}{\mathrm{}\mathrm{m}\tau }}\mathrm{}\mathrm{e}(\tau e_{(3)})i(\tau ^{}n_{(3)}^{IJ}\tau \stackrel{~}{n}_{(3)}^{IJ})\right\}.`$ (4.45) $``$ $`=`$ $`|M|^2\left[|M|^2e_{(3)}+2iN_{KL}\stackrel{~}{n}_{(3)}^{KL}\right].`$ (4.46) We are now ready to check whether these equations satisfy the relations expressed in Eqs. (3.10-3.12). It is immediate to see that they do if the following conditions are satisfied off-shell: $`e_{ij}`$ $`=`$ $`0,`$ (4.47) $`|M|^2\mathrm{}\mathrm{e}(e_{(3)})2\mathrm{}\mathrm{m}(N_{IJ}n_{(3)}^{IJ})`$ $`=`$ $`0.`$ (4.48) The first equation determines the 3-dimensional matric $`\gamma `$ as a function of the scalars $`N^{IJ}`$ and says that $`\gamma `$ is Ricci-flat is the projection $`𝒥^I_J`$ is constant. The second equation can be rewritten in the form $$_{\underset{¯}{i}}\left(\frac{Q^{\underset{¯}{i}}\xi ^{\underset{¯}{i}}}{|M|^2}\right)=0,$$ (4.49) and is the integrability condition of Eq. (4.39) for the 1-form $`\omega `$, whose existence we have assumed throughout all this analysis. Thus, it is not so much a necessary condition for supersymmetry as it is a necessary condition for the whole problem to be well defined. Let us summarize the results of this section: we have seen that, in the timelike case at hands, field configurations with a metric of the form Eq. (4.33), vector field strengths of the form Eq. (4.1) and any complex scalar $`\tau `$, and satisfying Eqs. (4.47) and (4.48) satisfy all the integrability conditions of the Killing spinor equations. On the other hand, all the equations of motion, including the Bianchi identities, are satisfied if the equations $$e_{(3)}^{}=0,n_{(3)}^{IJ}=0,e_{ij}=0,$$ (4.50) (were $`e_{(3)}^{}`$ and $`n_{(3)}^{IJ}`$ are defined in Eq. (4.37) and Eq. (4.36)) are satisfied, and automatically the integrability conditions are also satisfied. We are now ready to check whether the Killing spinor equations always admit solutions for those field configurations. Thus will help us in solving the integrability conditions Eqs. (4.47) and (4.48). ### 4.3 Solving the Killing spinor equations We have already checked that the equation $`\delta _ϵ\chi _I=0`$ is automatically solved by our field configurations, and we only have to check that the equations $`\delta _ϵ\psi _{aI}=0`$ can also be solved for them. Let us consider the timelike component first. It can be put in this form: $$|M|^1\left\{_tϵ_I\frac{1}{2}M^{KL}𝒟_iM_{KL}\gamma _{0i}\left[ϵ_I+i\sqrt{2}\gamma _0\frac{M_{IJ}}{|M|}ϵ^J\right]+\frac{i}{\sqrt{2}}\right|M|\stackrel{~}{𝒥}^K{}_{I}{}^{}𝒟_{i}^{}M_{KJ}\gamma _iϵ^J\}=0.$$ (4.51) Using the time-independent projector $`𝒥^I_J`$ we can split this equation into two equations: $`_tϵ_I\frac{1}{2}M^{KL}𝒟_iM_{KL}\gamma _{0i}\left[ϵ_I+i\sqrt{2}\gamma _0{\displaystyle \frac{M_{IJ}}{|M|}}ϵ^J\right]`$ $`=`$ $`0,`$ (4.52) $`\stackrel{~}{𝒥}^K{}_{I}{}^{}𝒟_{i}^{}M_{KJ}\gamma _iϵ^J`$ $`=`$ $`0.`$ (4.53) The first equation is solved by a time independent spinor because $$ϵ_I+i\sqrt{2}\gamma _0\frac{M_{IJ}}{|M|}ϵ^J=0,$$ (4.54) due to the Fierz identity $$M_{IJ}ϵ^J=\frac{i}{2}V^a\gamma _aϵ_I,$$ (4.55) and our choice of Vierbeins. For generic (i.e. not built from already-known Killing spinors) scalars $`M_{IJ}`$ the above relation would be a constraint breaking $`1/2`$ of the supersymmetries to be imposed on the Killing spinors whenever $`M^{KL}𝒟_iM_{KL}0`$. The counting of unbroken supersymmetries is, however, a bit more subtle and depends on the triviality of the $`U(1)`$ connection $`\xi `$: if $`\xi `$ is a total derivative the projection $`𝒥^I_J`$ is constant and a global $`SU(4)`$ rotation suffices to set to zero two of the chiral Killing spinors. This is the procedure followed by Tod in Ref. , where he solved the constant $`𝒥^I_J`$ (internally rigid) case by setting to zero two of the spinors, breaking the explicit $`SU(4)`$ covariance of the solutions. The solutions found by Tod preserve, then, generically, $`1/4`$ of the supersymmetries<sup>11</sup><sup>11</sup>11The conditions under which $`1/2`$ of the supersymmetries are preserved were studied in Ref. .. If $`𝒥^I_J`$ is not constant, $`\xi `$ is non-trivial and the 4 Killing spinors cannot be related by global $`SU(4)`$ rotations, but we are now going to see that this case can also be solved introducing a new projection on the Killing spinors which also reduces the amount of generically preserved supersymmetries to $`1/4`$. Now, using time-independence of the Killing spinors and Eq. (4.54), the spacelike components of $`\delta _ϵ\psi _{aI}=0`$ take the form $$\left[_i\frac{1}{2}\frac{M^{KL}_iM_{KL}}{|M|^2}\right]ϵ_I=0,$$ (4.56) which can be rewritten in the form $$(_ii\xi _i)(|M|^{1/2}ϵ_I)=0.$$ (4.57) The integrability condition for this equation is $$[R_{ijkl}\gamma ^{kl}+4i(d\xi )_{ij}]ϵ_I=0.$$ (4.58) This equation can be solved in essentially one way, up to local Lorentz transformations: $$R_{12}{}_{}{}^{12}=\pm 2(d\xi )_{12},\frac{1}{\sqrt{2}}(1i\gamma _{12})ϵ_I=0,$$ (4.59) the remaining components of the curvatures being zero. In terms of the connections we should have, in the appropriate Lorentz frame, the following relation between the 3-dimensional spin connection $`o^{ij}`$ and the $`U(1)`$ connection $`\xi `$: $$\xi =\pm \frac{1}{2}o^{12}(x^1,x^2)+\frac{1}{2}d\lambda (x^1,x^2,x^3),$$ (4.60) for some 3-dimensional 1-form $`\zeta `$ and some real scalar function $`\lambda `$. If complex scalars $`M^{IJ}`$ and 3-dimensional metric $`\gamma _{\underset{¯}{i}\underset{¯}{j}}`$ exist such that the above condition is met, then there are Killing spinors of the form $$ϵ_I=e^{\frac{i}{2}\lambda }|M|^{1/2}ϵ_I,(\xi \frac{1}{2}d\lambda )\frac{1}{\sqrt{2}}(1i\gamma _{12})ϵ_I=0.$$ (4.61) The relation between the spin connection and the $`U(1)`$ connection is just the requirement that the 3-dimensional metric has $`U(1)`$ holonomy, which implies that it is reducible to the direct product of a 2- and a 1-dimensional metric and, thus, can always be written in the form $$\gamma _{\underset{¯}{i}\underset{¯}{j}}dx^{\underset{¯}{i}}dx^{\underset{¯}{j}}=dx^2+2e^{2U(z,z^{})}dzdz^{},$$ (4.62) which, in turn, implies that $`\xi `$ is given by $$\xi =\pm \frac{i}{2}(_{\underset{¯}{z}}Udz_{\underset{¯}{z}^{}}Udz^{})+\frac{1}{2}d\lambda (x,z,z^{}).$$ (4.63) Let us summarize the results of this section. We have found that, to construct a supersymmetric configuration (not necessarily a solution) of pure, ungauged, $`N=4,d=4`$ supergravity amounts, now, to 1. Find a set of time-independent complex scalars $`M^{IJ}`$ satisfying $`\epsilon ^{IJKL}M_{IJ}M_{KL}=0`$ such that the $`U(1)`$ connection $`\xi `$ defined in Eq. (4.11) can be written in the form Eq. (4.63). The integrability condition Eq. (4.47) should automatically be solved by this choice. 2. Find $`\tau `$ by solving the integrability condition Eq. (4.49) of the defining equation of the 1-form $`\omega `$ (4.35). If we want the supersymmetric configuration to be a solution of the equations of motion, we also need to impose Eqs. (4.36) and (4.37), but we do not need to check the integrability condition Eq. (4.49). In the next section we study different solutions to these equations. ### 4.4 Supersymmetric configurations and solutions According to the recipe of the previous section, our first step in finding supersymmetric configurations and solutions is to find the complex scalars $`M^{IJ}`$ satisfying $`\epsilon ^{IJKL}M_{IJ}M_{KL}=0`$ and such that $`\xi `$ can be written in the form Eq. (4.63). The first condition can be easily met, for instance, by taking only $`M_{12},M_{13}`$ and $`M_{23}`$ non-vanishing, but we prefer not to make any specific choice that would break $`SU(4(`$ covariance. The second condition can be solved by the following Ansatz $$M_{IJ}=e^{i\lambda (x,z,z^{})}M(x,z,z^{})k_{IJ}(z),M=M^{},\lambda =\lambda ^{},\epsilon ^{IJKL}k_{IJ}k_{KL}=0,$$ (4.64) which give a connection $`\xi `$ of the form Eq. (4.63) with $$U=+\mathrm{ln}|k|,|k|^2k^{IJ}(z^{})k_{IJ}(z),$$ (4.65) and satisfies automatically the integrability condition Eq. (4.47). Solving the integrability condition Eq. (4.49) is considerably more difficult and considering solutions (instead of general configurations) simplifies the problem. We have found three families of solutions. 1. If the $`k_{IJ}`$ are constants, then, normalizing $`|k|^2=1`$ for simplicity, $`\xi =\frac{1}{2}d\lambda `$ and $`U=0`$. This is the case considered by Tod in Ref. and studied in detail in Ref. . Tod took advantage of the fact that $`d\xi =0`$ implies that $`𝒥^I_J`$ is constant and a global $`SU(4)`$ rotation can be used to set to zero two of the $`ϵ_I`$s. We will not do so, as this breaks the explicit $`SU(4)`$ covariance, but our results are, of course, equivalent. Eq. (4.36) takes the form $$_{\underset{¯}{i}}_{\underset{¯}{i}}_1=0,_1[(\mathrm{}\mathrm{m}\tau )^{1/2}e^{i\lambda }M]^1,$$ (4.66) and is solved by any arbitrary complex harmonic function $`_1`$. Using the above equation, Eq. (4.37) takes the form $$_{\underset{¯}{i}}_{\underset{¯}{i}}(_1\tau )=0,$$ (4.67) which is solved by $$\tau =_1/_2,_{\underset{¯}{i}}_2=0,$$ (4.68) another arbitrary complex harmonic function. The pair of harmonic functions and the constants determine completely the solutions. In particular $$|M|^2=M^2=\mathrm{}\mathrm{m}(\overline{_2}_1).$$ (4.69) 2. If $`e^{i\lambda }=M=1`$, the integrability condition Eq. (4.49) can be solved by taking $`\tau `$ constant. The only non-trivial equation of motion, Eq. (4.36) is solved using the holomorphicity of the $`k_{IJ}`$s. The metric takes the form $$ds^2=|k|^2(dt+\omega _{\underset{¯}{x}}dx)|k|^2dx^22dzdz^{},$$ (4.70) where $`\omega _{\underset{¯}{x}}`$ satisfies $$_{\underset{¯}{z}}\omega _{\underset{¯}{x}}_{\underset{¯}{x}}\omega _{\underset{¯}{z}}=_{\underset{¯}{z}^{}}|k|^2,_{\underset{¯}{z}^{}}\omega _{\underset{¯}{x}}_{\underset{¯}{x}}\omega _{\underset{¯}{z}^{}}=_{\underset{¯}{z}}|k|^2,_{\underset{¯}{z}^{}}\omega _{\underset{¯}{z}}_{\underset{¯}{z}}\omega _{\underset{¯}{z}^{}}=0.$$ (4.71) The metric and the supersymmetry projectors indicate that these solutions describe stationary strings lying along the coordinate $`x`$, in spite of the trivial axion field, which is the dual of the Kalb-Ramond 2-form $`B`$ that couples to strings. Observe, however, that the duality relation is not simply $`dB={}_{}{}^{}da`$: there are terms quadratic in the field strengths involved in the duality which must render $`B`$ non-trivial. The metric the the vector fields involved depends strongly on the choice of holomorphic $`k_{IJ}`$s. It is instructive to have an example completely worked out. Let us consider the simplest case: only $`k_{12}=\frac{1}{\sqrt{2}z}`$ non-trivial. This allows us to set $`\omega _{\underset{¯}{z}}=\omega _{\underset{¯}{z}^{}}=0`$. Then, $`|k|^2=|z|^2`$ and $`\omega _{\underset{¯}{x}}=2\mathrm{}\mathrm{e}(z^2)`$ and the full solution is given by $$\begin{array}{ccc}\hfill ds^2& =& \frac{1}{|z|^2}[dt+2\mathrm{}\mathrm{e}(z^2)dx]^2|z|^2dx^22dzdz^{},\hfill \\ & & \\ \hfill F_{12}& =& \frac{\sqrt{2}e^{\varphi _0/2}}{z^2}\{[dt+2\mathrm{}\mathrm{e}(z^2)dx]dzi{}_{}{}^{}[[dt+2\mathrm{}\mathrm{e}(z^2)dx]dz]\}=(F_{34})^{},\hfill \\ & & \\ \hfill \tau & =& \tau _0.\hfill \end{array}$$ (4.72) 3. The only solutions that we have found with $`\lambda `$ and the $`k_{IJ}(z)`$s simultaneously nontrivial have just $`\lambda =\lambda (x)`$ and $`M=M(x)`$ and are a superposition of the solutions with constant $`k_{IJ}`$ and the solutions with constant $`\lambda `$ in which these functions depend only on mutually transversal directions. Thus, these solutions depend on holomorphic functions $`k_{IJ}(z)`$ chosen with the same criteria as in the previous case, and a pair of complex functions $`_1,_2`$ linear in $`x`$ such that $`\mathrm{}\mathrm{m}\tau >0`$, and the metric is given by $$ds^2=(M|k|)^2(dt+\omega _{\underset{¯}{x}}dx)(M|k|)^2dx^22M^2dzdz^{},$$ (4.73) where $`M`$ is again given by Eq. (4.69). ## 5 The null case As we have mentioned before, the null case was completely solved by Tod in Ref. , but we include it her for the sake of completeness. As explained in Appendix B, in the null case all the spinors a proportional $`ϵ_I=\varphi _Iϵ`$. In the $`N=4,d=4`$ case at hands, $`ϵ_I`$ has a $`U(1)`$ charge under $`SL(2,)`$ transformations that has to be distributed between $`\varphi _I`$ and $`ϵ`$. We choose to have the $`\varphi _I`$ uncharged. Had we chosen to have $`\varphi _I`$ is charged with charge $`q_\varphi 0`$, then the real 1-form $$\zeta i\varphi _Id\varphi ^I,$$ (5.1) would transform as a $`U(1)`$ connection under $`SL(2,)`$ transformations as well and would play a role analogous to that of the connection $`\xi `$ in the timelike case. With our choice, $`\zeta `$ is just a $`U(1)`$ connection under the transformations Eq. (B.28) and covariantizes with respect to them the expressions that involve $`ϵ`$. We are now going to substitute $`ϵ_I=\varphi _Iϵ`$ into the KSEs and we are going to use the normalization condition to split the KSEs into three algebraic and one differential equation for $`ϵ`$. One of the algebraic equations for $`ϵ`$ will be a differential equation for $`\varphi _I`$. The substitution yields immediately $`𝒟_\mu \varphi _Iϵ+\varphi _I𝒟_\mu ϵ\frac{i}{2\sqrt{2}}(\mathrm{}\mathrm{m}\tau )^{1/2}F_{IJ}{}_{}{}^{+}{}_{\mu \nu }{}^{}\varphi _{}^{J}\gamma ^\nu ϵ^{}`$ $`=`$ $`0,`$ (5.2) $`\varphi _I{\displaystyle \frac{\overline{)}\tau }{\mathrm{}\mathrm{m}\tau }}ϵ\frac{1}{2\sqrt{2}}(\mathrm{}\mathrm{m}\tau )^{1/2}\overline{)}F_{IJ}{}_{}{}^{}\varphi _{}^{J}ϵ^{}`$ $`=`$ $`0.`$ (5.3) Acting on Eq. (5.2) with $`\varphi ^I`$ leads to $$𝒟_\mu ϵ=\varphi ^I𝒟_\mu \varphi _Iϵ,$$ (5.4) which takes the form $$\stackrel{~}{𝒟}_\mu ϵ(𝒟_\mu +i\zeta _\mu )ϵ=0,$$ (5.5) and becomes the only differential equation for $`ϵ`$. We have defined the derivative $`\stackrel{~}{𝒟}`$ covariant with respect to $`SL(2,)`$ and $`U(1)`$ local rotations under which $`ϵ`$ and $`\varphi _I`$ have charges $`+1`$ and $`1`$, respectively. Using Eq. (5.5) into Eq. (5.2) to eliminate $`𝒟_\mu ϵ`$ we obtain $$\stackrel{~}{𝒟}\varphi _Iϵ\frac{i}{2\sqrt{2}}(\mathrm{}\mathrm{m}\tau )^{1/2}F_{IJ}{}_{}{}^{+}{}_{\mu \nu }{}^{}\varphi _{}^{J}\gamma ^\nu ϵ^{}=0,$$ (5.6) which is one of the algebraic constraints for $`ϵ`$ and is a differential equation for $`\varphi _I`$. Acting with $`\varphi ^I`$ on Eq. (5.3) we see that it splits into two algebraic constraints for $`ϵ`$: $`\overline{)}\tau ϵ`$ $`=`$ $`0,`$ (5.7) $`\overline{)}F_{IJ}{}_{}{}^{}\varphi _{}^{J}ϵ^{}`$ $`=`$ $`0.`$ (5.8) Finally, we add to the system an auxiliary spinor $`\eta `$, introduced in Appendix B, with charges opposite to those of $`ϵ`$. The normalization condition Eq. (B.27) will be preserved if and only if $`\eta `$ satisfies a differential equation of the form $$\stackrel{~}{D}_\mu \eta +a_\mu ϵ=0,$$ (5.9) where $`a_\mu `$ is, in principle, an arbitrary vector with the right charges that transforms under the redefinitions Eqs. (B.36) and (B.37) as a connection $$a_\mu ^{}=a_\mu +_\mu \delta .$$ (5.10) In practice, however, $`a_\mu `$ cannot be completely arbitrary since the integrability conditions of the differential equation of $`\eta `$ have to be compatible with those of the differential equation for $`ϵ`$ and this requirement will determine $`a_\mu `$. Before we start a systematic analysis of these equations, it is worth comparing Eq. (5.5) to Eq. (4.57) and their integrability conditions which have the same structure except for the important detail of the dimensionality and signature. Therefore, we expect two main types of solutions: configurations with $`U(1)`$ holonomy on a 2-dimensional (spacelike) subspace and configurations with $`U(1)`$ holonomy in a null direction, which is the new possibility allowed by the Lorentzian signature. These expectations are also supported by the Fierz identities $`\overline{)}mϵ`$ $`=`$ $`iϵ,`$ (5.11) $`\mathit{}ϵ^{}`$ $`=`$ $`0,`$ (5.12) which are satisfied automatically here, but will be interpreted as projections. We will call these two possibilities $`B`$ and $`A`$ respectively. ### 5.1 Killing equations for the vector bilinears and first consequences We are now ready to derive equations involving the bilinears, in particular the vector bilinears which we construct with $`ϵ`$ and the auxiliary spinor $`\eta `$ introduced in Appendix B. First we deal with the equations that do not involve derivative of the spinors. Acting with $`\overline{ϵ}`$ on Eq. (5.6) and with $`\overline{ϵ^{}}\gamma ^\mu `$ on the complex conjugate of Eq. (5.8) we get $`\varphi ^IF_{IJ}{}_{}{}^{+}{}_{\mu \nu }{}^{}l_{}^{\nu }`$ $`=`$ $`0,`$ (5.13) $`ϵ^{IJKL}\varphi _JF_{KL}{}_{}{}^{+}{}_{\mu \nu }{}^{}l_{}^{\nu }`$ $`=`$ $`0.`$ (5.14) Acting with $`\overline{ϵ^{}}`$ and $`\overline{\eta ^{}}`$ on Eq. (5.7) we get<sup>12</sup><sup>12</sup>12The first of these equations had already been obtained in the general case Eq. (3.21). $`l\tau `$ $`=`$ $`0,`$ (5.15) $`m^{}\tau `$ $`=`$ $`0.`$ (5.16) Now, from Eqs. (5.5) and (5.9) we find $`_\mu l_\nu `$ $`=`$ $`0,`$ (5.17) $`\stackrel{~}{𝒟}_\mu n_\nu `$ $`=`$ $`a_\mu ^{}m_\nu a_\mu m_\nu ^{},`$ (5.18) $`\stackrel{~}{𝒟}_\mu m_\nu `$ $`=`$ $`a_\mu l_\nu .`$ (5.19) Let us now find the simplest implications of these equations. To start with,Eqs. (5.13) and (5.14), together, imply for nonvanishing $`\varphi _I`$<sup>13</sup><sup>13</sup>13This equation also follows from the general result Eq. (4.1) for vanishing scalars $`M_{IJ}`$. $$F_{IJ}{}_{}{}^{+}{}_{\mu \nu }{}^{}l_{}^{\nu }=0.$$ (5.20) Using Eq. (A.19), we see that the vector field strengths must take the form $`F_{IJ}^+`$ $`=`$ $`\frac{1}{2}_{IJ}lm^{},`$ (5.21) $`F_{IJ}^{}`$ $`=`$ $`\frac{1}{2}\stackrel{~}{}_{IJ}lm,`$ (5.22) where $`_{IJ}`$ is a skew-symmetric $`SU(4)`$ matrix of scalars to be determined and $`\stackrel{~}{}_{IJ}`$ is its $`SU(4)`$ dual. This solves completely Eq. (5.8), as can be seen using the Fierz identity $$l_\mu \gamma ^{\mu \nu }ϵ^{}=3l^\nu ϵ^{},$$ (5.23) and we can substitute Eq. (5.21) into Eq. (5.6) the only remaining equation in which vector field strengths occur. Using the Fierz identities $`\mathit{}ϵ^{}`$ $`=`$ $`0,`$ (5.24) $`\overline{)}m^{}ϵ^{}`$ $`=`$ $`iϵ,`$ (5.25) it takes the form $$\stackrel{~}{𝒟}_\mu \varphi _I\frac{1}{4\sqrt{2}}(\mathrm{}\mathrm{m}\tau )^{1/2}_{IJ}\varphi ^Jl_\mu =0,$$ (5.26) from which we find $$_{IJ}\varphi ^J=\frac{4\sqrt{2}}{(\mathrm{}\mathrm{m}\tau )^{1/2}}n^\mu \stackrel{~}{𝒟}_\mu \varphi _I.$$ (5.27) On the other hand, from Eqs. (5.15) and (5.16) we find that $$d\tau =A\widehat{l}+B\widehat{m}^{}.$$ (5.28) There are two cases to be considered here: case $`A`$ ($`B=0`$) and case $`B`$ ($`B0`$). In case $`B`$, we can write $$d\tau =B\left(\widehat{m}^{}+\frac{A}{B}\widehat{l}\right)=B\widehat{m}^{},$$ (5.29) after a redefinition of the type Eqs. (B.36) and (B.37). All the equations that we have written so far are covariant with respect to this kind of transformations and we just have to add primes (which we suppress immediately afterwards) everywhere. Thus, the case $`B`$ is equivalent to $`A=0`$ and we can always assume that either $`A`$ or $`B`$ is always zero. Since the connection $`Q`$ depends on $`\tau `$, the holonomy is different in these two cases. These are the two cases we mentioned at the end of the previous section and we will deal with them separately afterwards. ### 5.2 Equations of motion and integrability constraints Although we have not yet discussed the form of the metric, we already have enough information to study the equations of motion and check whether they satisfy the integrability conditions Eqs. (3.13)-(3.15). Using the results of the previous section, we can write the equations of motion in the form<sup>14</sup><sup>14</sup>14We have ignored all the terms that contain products $`AB`$ etc. $`_{\mu \nu }\frac{1}{2}g_{\mu \nu }^\rho _\rho `$ $`=`$ $`R_{\mu \nu }+\left[{\displaystyle \frac{|A|^2}{2(\mathrm{}\mathrm{m}\tau )^2}}+\frac{1}{16}\mathrm{}\mathrm{m}\tau ^2\right]l_\mu l_\nu +{\displaystyle \frac{|B|^2}{2(\mathrm{}\mathrm{m}\tau )^2}}m_{(\mu }m_{\nu )}^{},`$ (5.30) $``$ $`=`$ $`{\displaystyle \frac{1}{\mathrm{}\mathrm{m}\tau }}\left[l^\mu _\mu A^{}B^{}l^\mu a_\mu +m^\mu _\mu B^{}+\frac{i}{4}{\displaystyle \frac{|B|^2}{\mathrm{}\mathrm{m}\tau }}\right],`$ (5.31) $`\widehat{}_{IJ}\tau ^{}\widehat{}_{IJ}`$ $`=`$ $`i(\mathrm{}\mathrm{m}\tau )d(_{IJ}\widehat{l}\widehat{m}^{}).`$ (5.32) Substituting into Eqs. (3.13)-(3.15) and operating, we get $`R_{\mu \nu }l^\nu `$ $`=`$ $`0,`$ (5.33) $`R_{\mu \nu }m^\nu {\displaystyle \frac{|B|^2}{4(\mathrm{}\mathrm{m}\tau )^2}}m_\mu `$ $`=`$ $`0,`$ (5.34) $`l^\mu _\mu A^{}B^{}l^\mu a_\mu +m^\mu \stackrel{~}{𝒟}_\mu B^{}+\frac{i}{4}{\displaystyle \frac{|B|^2}{\mathrm{}\mathrm{m}\tau }}`$ $`=`$ $`0,`$ (5.35) $`B^{}_{IJ}\varphi ^J`$ $`=`$ $`0.`$ (5.36) We do not have a metric yet, but we can find $`R_{\mu \nu }l^\nu `$ and $`R_{\mu \nu }m^\nu `$ from the integrability conditions of Eqs. (5.5) and (5.9). Commuting the derivative and projecting with gamma matrices and spinors in the usual way, it is easy to find from Eq. (5.5) $`R_{\mu \nu }l^\nu `$ $`=`$ $`2i(d\zeta )_{\mu \nu }l^\nu ,`$ (5.37) $`R_{\mu \nu }m^\nu `$ $`=`$ $`+2i(d\zeta )_{\mu \nu }m^\nu 2i(dQ)_{\mu \nu }m^\nu `$ (5.38) $`=`$ $`+2i(d\zeta )_{\mu \nu }m^\nu +{\displaystyle \frac{|B|^2}{4(\mathrm{}\mathrm{m}\tau )^2}}m_\mu ,`$ and from Eq. (5.9) $`R_{\mu \nu }m^\nu `$ $`=`$ $`2i(d\zeta )_{\mu \nu }m^\nu 2i(dQ)_{\mu \nu }m^\nu +2(da)_{\mu \nu }l^\nu `$ (5.39) $`=`$ $`+2i(d\zeta )_{\mu \nu }m^\nu +{\displaystyle \frac{|B|^2}{4(\mathrm{}\mathrm{m}\tau )^2}}m_\mu +2(da)_{\mu \nu }l^\nu ,`$ $`R_{\mu \nu }n^\nu `$ $`=`$ $`2i(d\zeta )_{\mu \nu }n^\nu 2i(dQ)_{\mu \nu }n^\nu +2(da)_{\mu \nu }m^\nu `$ (5.40) $`=`$ $`2i(d\zeta )_{\mu \nu }n^\nu +2(da)_{\mu \nu }m^\nu .`$ Comparing now these three sets of equations, we get $$(d\zeta )_{\mu \nu }l^\nu =(d\zeta )_{\mu \nu }m^\nu =0,d\zeta =0,\zeta =d\alpha ,$$ (5.41) locally, and, eliminating $`\zeta `$ by a local phase redefinition, $`(da)_{\mu \nu }l^\nu `$ $`=`$ $`0,`$ (5.42) $`(da)_{\mu \nu }m^\nu `$ $`=`$ $`\frac{1}{2}R_{\mu \nu }n^\nu ,`$ (5.43) which tell us that $$da=\frac{1}{2}R_{z^{}u}\widehat{m}\widehat{m}^{}+\frac{1}{2}R_{uu}\widehat{l}\widehat{m}+C\widehat{l}\widehat{m}^{},$$ (5.44) where $`C`$ is a function to be chosen so as to make this equation (and, hence, Eq. (5.9)) integrable. Once $`\zeta `$ has been eliminated, we can solve Eq. (5.27) of $`_{IJ}`$ as follows: $$_{IJ}=\frac{8\sqrt{2}}{(\mathrm{}\mathrm{m}\tau )^{1/2}}n^\mu (_\mu \varphi _{[I})\varphi _{J]}.$$ (5.45) ### 5.3 Metric At this point we need information about the exact form of the metric. The most important piece of information comes from the covariant constancy of the null vector $`l^\mu `$. Metrics admitting a covariantly constant null vector are known as $`pp`$-wave metrics and were first described by Brinkmann in Refs. . Since $`l^\mu `$ is a Killing vector and $`d\widehat{l}=0`$ we can introduce the coordinates $`u`$ and $`v`$ $`l_\mu dx^\mu `$ $``$ $`du,`$ (5.46) $`l^\mu _\mu `$ $``$ $`{\displaystyle \frac{}{v}}.`$ (5.47) The previous results imply that all the objects we are dealing with ($`\tau ,\varphi _I,_{IJ}`$) are independent of $`v`$. Using these coordinates, a 4-dimensional $`pp`$-wave metric takes the form<sup>15</sup><sup>15</sup>15The components of the connection and the Ricci tensor of this metric can be found in Appendix D. $$ds^2=2du(dv+Kdu+\omega )2e^{2U}dzdz^{},\omega =\omega _{\underset{¯}{z}}dz+\omega _{\underset{¯}{z}^{}}dz^{},$$ (5.48) where all the functions in the metric are independent of $`v`$ and where either $`K`$ or the 1-form $`\omega `$ could, in principle, be removed by a coordinate transformation. In this case, however, we have to be very careful because we have already used part of the freedom we had to redefine the spinors, and, therefore, the null tetrad, and we have to check that the tetrad integrability equations (5.17)-(5.19) are satisfied by our choices of $`e^U,K`$ and $`\omega `$. We are now ready to study and solve each case separately. ### 5.4 Case $`A`$ This is the $`B=0`$ case. $`d\tau =A\widehat{l}`$ implies that $`\tau =\tau (u)`$ and $`A=\dot{\tau }`$. The connection $`Q`$ can be integrated $$Q=d\beta (u),$$ (5.49) and can be eliminated from all the equations by absorbing a phase into the spinors: $$e^{i\beta }ϵ=ϵ^{},e^{i\beta }\eta =\eta ^{},$$ (5.50) and similarly on the null tetrad. To fix the form of the metric, we study the antisymmetric part of Eq. (5.19) $$d\widehat{m}+\widehat{a}\widehat{l}=dU\widehat{m}+\widehat{a}\widehat{l}=0,$$ (5.51) which implies that $`U`$ only depends on $`u`$ and $$\widehat{a}=\dot{U}\widehat{m}+C\widehat{l},$$ (5.52) where $`D`$ is a function to be found. Substituting into the antisymmetric part of Eq. (5.18) we find $$d\widehat{n}+\widehat{a}^{}\widehat{m}+\widehat{a}\widehat{m}^{}=d\widehat{n}+C^{}\widehat{l}\widehat{m}+C\widehat{l}\widehat{m}^{}=0,$$ (5.53) which is solved by $$n=dv+Kdu,C^{}=e^U_{\underset{¯}{z}}K.$$ (5.54) Now, comparing Eq. (5.52) with Eq. (5.44) we find that $`R_{uz}=0`$ which implies (since $`\omega =0`$) that $`\dot{U}=0`$. Finally, to ensure supersymmetry, the integrability conditions Eqs. (5.33)-(5.36) have to be satisfied, and, with constant $`U`$ all of them are automatically satisfied. It also follows form the previous equations that the $`\varphi _I`$s can only depend on $`u`$ and $`_{IJ}`$ is given by $$_{IJ}=\frac{8\sqrt{2}}{(\mathrm{}\mathrm{m}\tau )^{1/2}}\dot{\varphi }_{[I}\varphi _{J]}.$$ (5.55) Now, let us consider the equations of motion. The scalar, Maxwell and Bianchi equations are automatically satisfied and the Einstein equation can be solved by a $`K`$ satisfying $$2_{\underset{¯}{z}}_{\underset{¯}{z}^{}}K=\frac{|\dot{\tau }|^2}{(\mathrm{}\mathrm{m}\tau )^2}+\frac{1}{16}\mathrm{}\mathrm{m}\tau ^2.$$ (5.56) These solutions preserve generically $`1/4`$ of the supersymmetries. ### 5.5 Case $`B`$ This is the $`A=0`$ case. If we choose $`m^{}=e^Udz^{}`$, then $`d\tau =Bm^{}`$ implies $`\tau =\tau (z^{})`$ and $`Be^U=_{\underset{¯}{z}^{}}\tau `$. Substituting the corresponding connection 1-form $`Q`$ into Eq. (5.19) one finds $`B^{}`$ $`=`$ $`{\displaystyle \frac{g(z,u)}{(\mathrm{}\mathrm{m}\tau )^{1/2}}},`$ (5.57) $`\widehat{a}`$ $`=`$ $`_{\underset{¯}{u}}\mathrm{ln}g\widehat{m}+D\widehat{l},`$ (5.58) where $`g`$ is a holomorphic function of $`z`$ and $`D`$ is a function to be determined. The first of these relations tells us that $$_{\underset{¯}{z}}\tau ^{}=\frac{e^U}{(\mathrm{}\mathrm{m}\tau )^{1/2}}g(z,u),$$ (5.59) is a holomorphic function of $`z`$, independent of $`u`$, and taking the derivative of both sides with respect to $`z^{}`$ we get $$\frac{e^U}{(\mathrm{}\mathrm{m}\tau )^{1/2}}=f(u),g(z,u)=\frac{h(z)}{f(u)},$$ (5.60) where $`f(u)`$ is a real function of $`u`$. Substituting now $`\widehat{a}`$ into the antisymmetric part of Eq. (5.18) we find that $`\widehat{n}`$ is given by $$\widehat{n}=dv+\omega ,$$ (5.61) (so $`K=0`$ in the metric Eq. (5.48)) where the 1-form $`\omega `$ satisfies $$f_{\underset{¯}{z}\underset{¯}{z}^{}}=e^{2U}_{\underset{¯}{u}}\mathrm{ln}(B/B^{})=0,$$ (5.62) and $`D`$ is given by $$D^{}=\dot{\omega }_{\underset{¯}{z}}e^U.$$ (5.63) Now that we have determined $`\widehat{a}`$ we have to check that it satisfies the integrability condition Eq. (5.44). This requires the following equations to be satisfied: $`R_{uz^{}}+\frac{i}{2}{\displaystyle \frac{_{\underset{¯}{u}}\mathrm{ln}fB}{\mathrm{}\mathrm{m}\tau }}`$ $`=`$ $`0,`$ (5.64) $`R_{uu}[_{\underset{¯}{u}}^2\mathrm{ln}f+_{\underset{¯}{u}}\mathrm{ln}f_{\underset{¯}{u}}\mathrm{ln}f]2e^U_{\underset{¯}{z}}D`$ $`=`$ $`0,`$ (5.65) $`Ce^U_{\underset{¯}{z}^{}}D`$ $`=`$ $`0.`$ (5.66) Comparing with the integrability conditions Eqs. (5.33)-(5.36), we conclude that $`f`$ must be a constant that we normalize $`f=1`$ and that $`\omega `$ must be exact, and we can eliminate it. Further, the $`\varphi _I`$s must be constant and the vector field strengths must vanish. All the equations of motion are automatically satisfied in these conditions, and the solutions are the stringy cosmic strings of Ref. . Our result differs from Tod’s, who used $`\tau `$ and $`\tau ^{}`$ as coordinates and found very similar solutions with nontrivial $`\omega `$ that depend in a very complicated way on a function $`g(\tau ,u)`$ an its complex conjugate. This function could be eliminated by a coordinate change in which all the $`u`$ dependence and the 1-form $`\omega `$ disappear, recovering the stringy cosmic string solutions. ## Acknowledgements The authors would like to thank P. Meessen for interesting and insightful conversations and A. Fernández del Río for her help in the early stages of this work. T.O. is indebted to M.M. Fernández for her support. This work has been supported in part by the Spanish grant BFM2003-01090. ## Appendix A Conventions ### A.1 Tensors We use Greek letters $`\mu ,\nu ,\rho ,\mathrm{}`$ as (curved) tensor indices in a coordinate basis and Latin letters $`a,b,c\mathrm{}`$ as (flat) tensor indices in a tetrad basis. Underlined indices are always curved indices. We symmetrize $`()`$ and antisymmetrize $`[]`$ with weight one (i.e. dividing by $`n!`$). We use mostly minus signature $`(+)`$. $`\eta `$ is the Minkowski metric and a general metric is denoted by $`g`$. Flat and curved indices are related by tetrads $`e_a^\mu `$ and their inverses $`e^a_\mu `$, satisfying $$e_a{}_{}{}^{\mu }e_{b}^{}{}_{}{}^{\nu }g_{\mu \nu }^{}=\eta _{ab},e^a{}_{\mu }{}^{}e_{}^{b}{}_{\nu }{}^{}\eta _{ab}^{}=g_{\mu \nu }.$$ (A.1) $``$ is the total (general- and Lorentz-) covariant derivative, whose action on tensors and spinors ($`\psi `$) is given by $$\begin{array}{ccc}\hfill _\mu \xi ^\nu & =& _\mu \xi ^\nu +\mathrm{\Gamma }_{\mu \rho }{}_{}{}^{\nu }\xi _{}^{\rho },\hfill \\ & & \\ \hfill _\mu \xi ^a& =& _\mu \xi ^a+\omega _{\mu b}{}_{}{}^{a}\xi _{}^{b},\hfill \\ & & \\ \hfill _\mu \psi & =& _\mu \psi \frac{1}{4}\omega _\mu {}_{}{}^{ab}\gamma _{ab}^{}\psi ,\hfill \end{array}$$ (A.2) where $`\gamma _{ab}`$ is the antisymmetric product of two gamma matrices (see next section), $`\omega _{\mu b}^a`$ is the spin connection and $`\mathrm{\Gamma }_{\mu \rho }^\nu `$ is the affine connection. The respective curvatures are defined through the Ricci identities $$\begin{array}{ccc}\hfill [_\mu ,_\nu ]\xi ^\rho & =& R_{\mu \nu \sigma }{}_{}{}^{\rho }(\mathrm{\Gamma })\xi ^\sigma +T_{\mu \nu }{}_{}{}^{\sigma }_{\sigma }^{}\xi ^\rho ,\hfill \\ & & \\ \hfill [_\mu ,_\nu ]\xi ^a& =& R_{\mu \nu b}{}_{}{}^{a}(\omega )\xi ^b,\hfill \\ & & \\ \hfill [_\mu ,_\nu ]\psi & =& \frac{1}{4}R_{\mu \nu }{}_{}{}^{ab}(\omega )\gamma _{ab}\psi .\hfill \end{array}$$ (A.3) and given in terms of the connections by $$\begin{array}{ccc}\hfill R_{\mu \nu \rho }{}_{}{}^{\sigma }(\mathrm{\Gamma })& =& 2_{[\mu }\mathrm{\Gamma }_{\nu ]\rho }{}_{}{}^{\sigma }+2\mathrm{\Gamma }_{[\mu |\lambda }{}_{}{}^{\sigma }\mathrm{\Gamma }_{\nu ]\rho }^{}{}_{}{}^{\lambda },\hfill \\ & & \\ \hfill R_{\mu \nu a}{}_{}{}^{b}(\omega )& =& 2_{[\mu }\omega _{\nu ]a}{}_{}{}^{b}2\omega _{[\mu |a}{}_{}{}^{c}\omega _{|\nu ]c}^{}{}_{}{}^{b}.\hfill \end{array}$$ (A.4) These two connections are related by the tetrad postulate $$_\mu e_a{}_{}{}^{\mu }=0,$$ (A.5) by $$\omega _{\mu a}{}_{}{}^{b}=\mathrm{\Gamma }_{\mu a}{}_{}{}^{b}+e_a{}_{}{}^{\nu }_{\mu }^{}e_\nu {}_{}{}^{b},$$ (A.6) which implies that the curvatures are, in turn, related by $$R_{\mu \nu \rho }{}_{}{}^{\sigma }(\mathrm{\Gamma })=e_\rho {}_{}{}^{a}e_{}^{\sigma }{}_{b}{}^{}R_{\mu \nu a}^{}{}_{}{}^{b}(\omega ).$$ (A.7) Finally, metric compatibility and torsionlessness fully determine the connections to be of the form $$\begin{array}{ccc}\hfill \mathrm{\Gamma }_{\mu \nu }^\rho & =& \frac{1}{2}g^{\rho \sigma }\left\{_\mu g_{\nu \sigma }+_\nu g_{\mu \sigma }_\sigma g_{\mu \nu }\right\},\hfill \\ & & \\ \hfill \omega _{abc}& =& \mathrm{\Omega }_{abc}+\mathrm{\Omega }_{bca}\mathrm{\Omega }_{cab},\mathrm{\Omega }_{ab}{}_{}{}^{c}=e_a{}_{}{}^{\mu }e_{b}^{}{}_{}{}^{\nu }_{[\mu }^{}e^c{}_{\nu ]}{}^{}.\hfill \end{array}$$ (A.8) The 4-dimensional fully antisymmetric tensor is defined in flat indices by tangent space by $$ϵ^{0123}=+1,ϵ_{013}=1,$$ (A.9) and in curved indices by $$ϵ^{\mu _1\mathrm{}\mu _3}=\sqrt{|g|}e^{\mu _1}{}_{a_1}{}^{}\mathrm{}e^{\mu _3}{}_{a_3}{}^{}ϵ_{}^{a_3\mathrm{}a_3},$$ (A.10) so, with upper indices, is independent of the metric and has the same value as with flat indices. We define the (Hodge) dual of a completely antisymmetric tensor of rank $`k`$, $`F_{(k)}`$ by $${}_{}{}^{}F_{(k)}^{}{}_{}{}^{\mu _1\mathrm{}\mu _{(dk)}}=\frac{1}{k!\sqrt{|g|}}ϵ^{\mu _1\mathrm{}\mu _{(dk)}\mu _{(dk+1)}\mathrm{}\mu _d}F_{(k)\mu _{(dk+1)}\mathrm{}\mu _d}.$$ (A.11) Differential forms of rank $`k`$ are normalized as follows: $$F_{(k)}\frac{1}{k!}F_{(k)}{}_{}{}^{\mu _1\mathrm{}\mu _k}dx^1\mathrm{}dx^k.$$ (A.12) For any 4-dimensional 2-form, we define $$F^\pm \frac{1}{2}(F\pm i{}_{}{}^{}F),\pm i{}_{}{}^{}F_{}^{\pm }=F^\pm .$$ (A.13) For any two 2-forms $`F,G`$, we have $$F^\pm G^{}=0,F^\pm {}_{[\mu }{}^{}{}_{}{}^{\rho }G^{}{}_{\nu ]\rho }{}^{}=0.$$ (A.14) Given any 2-form $`F=\frac{1}{2}F_{\mu \nu }dx^\mu dx^\nu `$ and a non-null 1-form $`\widehat{V}=V_\mu dx^\mu `$, we can express $`F`$ in the form $$F=V^2[E\widehat{V}{}_{}{}^{}(B\widehat{V})],E_\mu F_{\mu \nu }V^\nu ,B_\mu {}_{}{}^{}F_{\mu \nu }^{}V^\nu .$$ (A.15) For the complex combinations $`F^\pm `$ we have $$F^\pm =V^2[C^\pm \widehat{V}\pm i{}_{}{}^{}(C^\pm \widehat{V})],C_\mu ^\pm F_{\mu \nu }^\pm V^\nu .$$ (A.16) If we have a (real) null vector $`l^\mu `$, we can always add three more null vectors $`n^\mu ,m^\mu ,m^\mu `$ to construct a complex null tetrad such that the local metric in this basis takes the form $$\left(\begin{array}{cccc}\hfill 0& \hfill 1& \hfill 0& \hfill 0\\ \hfill 1& \hfill 0& \hfill 0& \hfill 0\\ \hfill 0& \hfill 0& \hfill 0& \hfill 1\\ \hfill 0& \hfill 0& \hfill 1& \hfill 0\end{array}\right)$$ (A.17) with the ordering $`(l,n,m,m^{})`$. For the local volume element we obtain $`ϵ^{lnmm^{}}=i`$. The general expansion in the dual basis of 1-forms $`(\widehat{l},\widehat{n},\widehat{m},\widehat{m}^{})`$ of $`F^+`$ depends on three arbitrary complex functions $`a,b,c`$ $$F^+=a\left(\widehat{l}\widehat{n}+\widehat{m}\widehat{m}^{}\right)+b\widehat{l}\widehat{m}^{}+c\widehat{n}\widehat{m},F^{}=(F^+)^{}.$$ (A.18) Then, in this case, $`F`$ is not completely determined by its contraction with the null vector $`l`$, but $$F^+=L^\pm \widehat{n}\pm {}_{}{}^{}(L^\pm \widehat{n})+b\widehat{l}\widehat{m},L_\mu ^\pm F^\pm {}_{\mu \nu }{}^{}l_{}^{\nu }=al_\mu cm_\mu .$$ (A.19) ### A.2 Gamma matrices and spinors We work with a purely imaginary representation $$\gamma ^a=\gamma ^a,$$ (A.20) and our convention for their anticommutator is $$\{\gamma ^a,\gamma ^b\}=+2\eta ^{ab}.$$ (A.21) Thus, $$\gamma ^0\gamma ^a\gamma ^0=\gamma ^a=\gamma ^{a1}=\gamma _a.$$ (A.22) The chirality matrix is defined by $$\gamma _5i\gamma ^0\gamma ^1\gamma ^2\gamma ^3=\frac{i}{4!}ϵ_{abcd}\gamma ^a\gamma ^b\gamma ^c\gamma ^d,$$ (A.23) and satisfies $$\gamma _5{}_{}{}^{}=\gamma _5{}_{}{}^{}=\gamma _5,(\gamma _5)^2=1.$$ (A.24) With this chirality matrix, we have the identity $$\gamma ^{a_1\mathrm{}a_n}=\frac{(1)^{\left[n/2\right]}i}{(4n)!}ϵ^{a_1\mathrm{}a_nb_1\mathrm{}b_{4n}}\gamma _{b_1\mathrm{}b_{4n}}\gamma _5.$$ (A.25) Our convention for Dirac conjugation is $$\overline{\psi }=i\psi ^{}\gamma _0.$$ (A.26) Using the identity Eq. (A.25) the general $`d=4`$ Fierz identity for commuting spinors takes the form $$\begin{array}{ccc}\hfill (\overline{\lambda }M\chi )(\overline{\psi }N\phi )& =& \frac{1}{4}(\overline{\lambda }MN\phi )(\overline{\psi }\chi )+\frac{1}{4}(\overline{\lambda }M\gamma ^aN\phi )(\overline{\psi }\gamma _a\chi )\frac{1}{8}(\overline{\lambda }M\gamma ^{ab}N\phi )(\overline{\psi }\gamma _{ab}\chi )\hfill \\ & & \\ & & \frac{1}{4}(\overline{\lambda }M\gamma ^a\gamma _5N\phi )(\overline{\psi }\gamma _a\gamma _5\chi )+\frac{1}{4}(\overline{\lambda }M\gamma _5N\phi )(\overline{\psi }\gamma _5\chi ).\hfill \end{array}$$ (A.27) We use 4-component chiral spinors whose chirality is related to the position of the $`SU(4)`$ index: $$\gamma _5\chi _I=+\chi _I,\gamma _5\psi _{\mu I}=\psi _{\mu I},\gamma _5ϵ_I=ϵ_I.$$ (A.28) Both (chirality and position of the $`SU(4)`$ index) are reversed under complex conjugation: $$\gamma _5\chi _I^{}\gamma _5\chi ^I=\chi ^I,\gamma _5\psi _{\mu I}^{}\gamma _5\psi _\mu {}_{}{}^{I}=+\psi _\mu {}_{}{}^{I},\gamma _5ϵ_I^{}\gamma _5ϵ^I=+ϵ^I.$$ (A.29) We take this fact into account when Dirac-conjugating chiral spinors: $$\overline{\chi }^Ii(\chi _I)^{}\gamma _0,\overline{\chi }^I\gamma _5=\overline{\chi }^I,\mathrm{etc}.$$ (A.30) The sum of the two chiral spinors related by complex conjugation gives a standard (real) Majorana spinor with an $`SU(4)`$ index with the complicated transformation rule of Ref. . ## Appendix B Fierz identities for bilinears Here we are going to work with an arbitrary number $`N`$ of chiral spinors, although we are ultimately interested in the $`N=4`$ case only. Whenever there are special results for particular values of $`N`$, we will explicitly say so. We should bear in mind that the maximal number of independent chiral spinors is 2 and, for $`N>2`$ (in particular for $`N=4`$) $`N`$ spinors cannot be linearly independent at a given point. This trivial fact has important consequences. Given $`N`$ chiral commuting spinors $`ϵ_I`$ and their complex conjugates $`ϵ^I`$ we can constructed the following bilinears that are not obviously related via Eq. (A.25): 1. A complex matrix of scalars $$M_{IJ}\overline{ϵ}_Iϵ_J,M^{IJ}\overline{ϵ}^Iϵ^J=(M_{IJ})^{},$$ (B.1) which is antisymmetric $`M_{IJ}=M_{JI}`$. 2. A complex matrix of vectors $$V^I{}_{Ja}{}^{}i\overline{ϵ}^I\gamma _aϵ_J,V_I{}_{}{}^{J}{}_{a}{}^{}i\overline{ϵ}_I\gamma _aϵ^J=(V^I{}_{Ja}{}^{})^{},$$ (B.2) which is Hermitean: $$(V^I{}_{Ja}{}^{})^{}=V_I{}_{}{}^{J}{}_{a}{}^{}=V^J{}_{Ia}{}^{}=(V^I{}_{Ja}{}^{})^T.$$ (B.3) 3. A complex matrix of 2-forms $$\mathrm{\Phi }_{IJab}\overline{ϵ}_I\gamma _{ab}ϵ_J,\mathrm{\Phi }^{IJ}{}_{ab}{}^{}\overline{ϵ}^I\gamma _{ab}ϵ^J=(M_{IJ})^{},$$ (B.4) which is symmetric in the $`SU(N)`$ indices $`\mathrm{\Phi }_{IJab}=\mathrm{\Phi }_{JIab}`$ and, further, $${}_{}{}^{}\mathrm{\Phi }_{IJab}^{}=i\mathrm{\Phi }_{IJab}\mathrm{\Phi }_{IJab}=\mathrm{\Phi }_{IJ}{}_{}{}^{+}{}_{ab}{}^{}.$$ (B.5) As we are going to see, this matrix of 2-forms can be expressed entirely in terms of the scalar and vector bilinears. It is straightforward to get identities for the products of these bilinears using the Fierz identity Eq. (A.27). First, the products of scalars: $`M_{IJ}M_{KL}`$ $`=`$ $`\frac{1}{2}M_{IL}M_{KJ}\frac{1}{8}\mathrm{\Phi }_{IL}\mathrm{\Phi }_{KJ},`$ (B.6) $`M_{IJ}M^{KL}`$ $`=`$ $`\frac{1}{2}V^L{}_{I}{}^{}V^K{}_{J}{}^{}.`$ (B.7) From Eq. (B.6) immediately follows $$M_{I[J}M_{KL]}=0,$$ (B.8) which is a particular case of the Fierz identity $$ϵ_{[J}M_{KL]}=0.$$ (B.9) For $`N=4,8,\mathrm{}`$, Eq. (B.8) implies, in turn $$\mathrm{Pf}M=0\mathrm{det}M=0.$$ (B.10) For $`N=4`$ we can define the $`SU(4)`$-dual of $`M_{IJ}`$ $$\stackrel{~}{M}_{IJ}\frac{1}{2}\epsilon _{IJKL}M^{KL},\epsilon ^{1234}=\epsilon _{1234}=+1,$$ (B.11) and the vanishing of the Pfaffian implies $$\stackrel{~}{M}_{IJ}M^{IJ}=0.$$ (B.12) From Eq. (B.7) and the antisymmetry of $`M`$ immediately follows $$V^I{}_{L}{}^{}V^K{}_{J}{}^{}=V^I{}_{J}{}^{}V^K{}_{L}{}^{}=V^K{}_{L}{}^{}V^I{}_{J}{}^{},$$ (B.13) which implies that all the vector bilinears $`V^I_{Ja}`$ are null: $$V^I{}_{J}{}^{}V^I{}_{J}{}^{}=0.$$ (B.14) On the other hand, from Eqs. (B.13) and (B.7) follows the real $`SU(N)`$-invariant combination of vectors $`V_aV^I_{Ia}`$ is always non-spacelike: $$V^2=V^I{}_{J}{}^{}V^J{}_{I}{}^{}=2M^{IJ}M_{IJ}0.$$ (B.15) The products of $`M`$ with the other bilinears<sup>16</sup><sup>16</sup>16We omit the product $`M_{IJ}\mathrm{\Phi }_{KLab}`$ which will not be used. give $`M_{IJ}V^K_{La}`$ $`=`$ $`\frac{1}{2}M_{IL}V^K{}_{Ja}{}^{}+\frac{1}{2}\mathrm{\Phi }_{ILba}V^K{}_{J}{}^{}{}_{}{}^{b},`$ (B.16) $`M_{IJ}\mathrm{\Phi }^{KL}_{ab}`$ $`=`$ $`V^L{}_{I[a|}{}^{}V_{}^{K}{}_{J|b]}{}^{}\frac{i}{2}ϵ_{ab}{}_{}{}^{cd}V_{}^{L}{}_{Ic}{}^{}V_{}^{K}{}_{Jd}{}^{}.`$ (B.17) Now, let us consider the product of two arbitrary vectors<sup>17</sup><sup>17</sup>17The product $`V^I{}_{Ja}{}^{}V_{L}^{}{}_{}{}^{K}_b`$ gives a different identity that will not be used: $$V^I{}_{Ja}{}^{}V_{}^{K}{}_{Lb}{}^{}=\frac{i}{2}ϵ_{ab}{}_{}{}^{cd}V_{}^{I}{}_{Lc}{}^{}V_{}^{K}{}_{Jd}{}^{}+V^I{}_{L(a|}{}^{}V_{}^{K}{}_{J|b)}{}^{}\frac{1}{2}g_{ab}V^I{}_{L}{}^{}V^K{}_{J}{}^{}.$$ (B.18) For $`V^2`$ this identity allows us to write the metric in the form $$g_{ab}=2V^2[V_aV_bV^I{}_{Ja}{}^{}V_{}^{J}{}_{Ib}{}^{}].$$ (B.19) Following Tod , for $`V^20`$ we introduce $$𝒥^I{}_{J}{}^{}\frac{2M^{IK}M_{JK}}{|M|^2}=\frac{2VV^I_J}{V^2},|M|^2M^{LM}M_{LM}=\frac{1}{2}V^2.$$ (B.20) Using Eq. (B.6) we can show that it is a Hermitean projector whose trace equals 2: $$𝒥^I{}_{J}{}^{}𝒥_{}^{J}{}_{K}{}^{}=𝒥^I{}_{K}{}^{},𝒥^I{}_{I}{}^{}=+2.$$ (B.21) Further, using the general Fierz identity we find $$𝒥^I{}_{J}{}^{}ϵ_{}^{J}=ϵ^I,ϵ_I𝒥^I{}_{J}{}^{}=ϵ_J,$$ (B.22) which should be understood for $`N>2`$ of the fact that the $`ϵ^I`$ are not linearly independent<sup>18</sup><sup>18</sup>18For $`N=2`$ $`𝒥^I{}_{J}{}^{}=\delta ^I_J`$. See later on.. As a consequence of the above identity, the contraction of $`𝒥`$ with any of the bilinears is the identity. Using this result and Eq. (B.17), we find $$\mathrm{\Phi }^{KL}{}_{ab}{}^{}=\frac{2M^{IK}M_{IJ}}{|M|^2}\mathrm{\Phi }^{JL}{}_{ab}{}^{}=\frac{2M^{IK}}{|M|^2}V^L{}_{I[a}{}^{}V_{b]}^{}i\frac{M^{IK}}{|M|^2}ϵ_{ab}{}_{}{}^{cd}V_{}^{L}{}_{Ic}{}^{}V_{d}^{}.$$ (B.23) Other useful identities are $$\frac{M_{IJ}M^{KL}}{|M|^2}=𝒥^K{}_{[I}{}^{}𝒥_{}^{L}{}_{J]}{}^{},$$ (B.24) and $$\frac{2\stackrel{~}{M}^{IK}\stackrel{~}{M}_{JK}}{|M|^2}=\delta ^I{}_{J}{}^{}𝒥^I{}_{J}{}^{}\stackrel{~}{𝒥}{}_{}{}^{I}{}_{J}{}^{},$$ (B.25) which is the complementary projector. In the null case $`V^2=|M|^2=0`$ it is customary to write $`l_aV^I_{Ia}`$. Since $`|M|^2`$ is a sum of positive numbers, each of them must vanish independently, i.e. $`M^{IJ}=0`$. This implies that all spinors $`ϵ^I`$ are proportional and one can write $$ϵ_I=\varphi _Iϵ,$$ (B.26) for some complex functions $`\varphi _I`$ which transform as an $`SU(4)`$ vector, and some negative-chirality spinor $`ϵ`$. These are defined up to a rescaling by a complex function and opposite weights. Part of this freedom can be fixed by normalizing $$\varphi _I\varphi ^I=1,\varphi ^I\varphi _I^{}.$$ (B.27) Then, the only freedom that remains in the definition of $`\varphi ^I`$ is a change by a local phase $`\theta (x)`$ $$\varphi _Ie^{i\theta }\varphi _I,ϵe^{i\theta }ϵ.$$ (B.28) In this case on can construct another Hermitean projector $`𝒦^I_J`$ that plays a role analogous to that of $`𝒥^I_J`$ in the non-null case: $$𝒦^I{}_{J}{}^{}\varphi ^I\varphi _J,$$ (B.29) which satisfies $$𝒦^I{}_{J}{}^{}𝒦_{}^{J}{}_{K}{}^{}=𝒦^I{}_{K}{}^{},𝒦^I{}_{I}{}^{}=+1,$$ (B.30) and $$𝒦^I{}_{J}{}^{}ϵ_{}^{J}=ϵ^I,ϵ_I𝒦^I{}_{J}{}^{}=ϵ_J,$$ (B.31) which expresses the known fact that only one spinor is linearly independent in this case. In the null case, all the vector bilinears are also proportional to the null vector $`l`$: $$V^I{}_{Ja}{}^{}=𝒦^I{}_{J}{}^{}l_{a}^{}.$$ (B.32) Once $`ϵ`$ is given, we may introduce an auxiliary spinor with the same chirality and opposite $`U(1)`$ charge as $`ϵ`$ and normalized against $`ϵ`$ by $$\overline{ϵ}\eta =\frac{1}{2},$$ (B.33) where $`\overline{ϵ}=iϵ^T\gamma _0`$. With both spinors we can construct a complex null tetrad with metric Eq. (A.17) as follows: $$l_\mu =i\overline{ϵ^{}}\gamma _\mu ϵ,n_\mu =i\overline{\eta ^{}}\gamma _\mu \eta ,m_\mu =i\overline{ϵ^{}}\gamma _\mu \eta =i\overline{\eta }\gamma _\mu ϵ^{},m_\mu ^{}=i\overline{ϵ}\gamma _\mu \eta ^{}=i\overline{\eta ^{}}\gamma _\mu ϵ.$$ (B.34) The normalization condition (B.27) does not fix completely the auxiliary spinor $`\eta `$ and the freedom in the choice of $`\eta `$ becomes a freedom in the null tetrad. First of all, there is a $`U(1)`$ freedom Eq. (B.28) under which $`\eta ^{}=e^{i\theta }\eta `$ and $$l^{}=l,n^{}=n,m^{}=e^{2i\theta }m.$$ (B.35) Further, we can also shift $`\eta `$ by terms proportional to $`ϵ`$ preserving the normalization $$\eta ^{}=\eta +\delta ϵ.$$ (B.36) Under this redefinition of $`\eta `$, the null tetrad transforms as follows: $$l^{}=l,n^{}=n+\delta ^{}m+\delta m^{}+|\delta |^2l,m^{}=m+\delta l.$$ (B.37) ### B.1 The $`N=2`$ case Here we describe some of the peculiarities of the $`N=2`$ case in which the number of spinors is precisely the necessary to construct a basis at each point. In the $`N=2`$ case there is only one independent (complex) scalar $`X`$ since $$\overline{ϵ}_Iϵ_J=Xϵ_{IJ},$$ (B.38) where $`ϵ_{IJ}`$ is the (constant) 2-dimensional totally antisymmetric tensor. It follows that $$|M|^2=2|X|^2,$$ (B.39) and, using $`ϵ_{IJ}ϵ^{KL}=\delta _{IJ}^{KL}`$ we can show that the projector $$𝒥^I{}_{J}{}^{}=\delta ^I{}_{J}{}^{}.$$ (B.40) In the $`|M|^20`$ case, the four vector bilinears $`V^I_{J\mu }`$ can be used as a null tetrad $$l_\mu =V^1{}_{1\mu }{}^{},n_\mu =V^2{}_{2\mu }{}^{},m_\mu =V^1{}_{2\mu }{}^{},m_\mu ^{}=V^2{}_{1\mu }{}^{},.$$ (B.41) Alternatively. one can use the four combinations $$V^a{}_{\mu }{}^{}\frac{1}{\sqrt{2}}V^I{}_{J\mu }{}^{}(\sigma ^a)_{}^{J}{}_{I}{}^{},$$ (B.42) with $`\sigma ^0=1`$ and $`\sigma ^i`$ the three (traceless, Hermitean) Pauli matrices as an orthonormal tetrad in which $`V^0`$ is timelike and the $`V^i`$ are spacelike. ## Appendix C Connection and curvature of the conformastationary metric A conformastationary metric has the general form $$ds^2=|M|^2(dt+\omega )^2|M|^2\gamma _{\underset{¯}{i}\underset{¯}{j}}dx^idx^j,i,j=1,2,3,$$ (C.1) where all components of the metric are independent of the time coordinate $`t`$. Choosing the Vielbein basis $$(e_\mu {}_{}{}^{a})=\left(\begin{array}{cc}|M|& |M|\omega _{\underset{¯}{i}}\\ & \\ 0& |M|^1v_{\underset{¯}{i}}^j\end{array}\right),(e_a{}_{}{}^{\mu })=\left(\begin{array}{cc}|M|^1& |M|\omega _i\\ & \\ 0& |M|v_i^{\underset{¯}{j}}\end{array}\right),$$ (C.2) where $$\gamma _{\underset{¯}{i}\underset{¯}{j}}=v_{\underset{¯}{i}}{}_{}{}^{k}v_{\underset{¯}{j}}^{}{}_{}{}^{l}\delta _{kl}^{},v_i{}_{}{}^{\underset{¯}{k}}v_{\underset{¯}{k}}^{}{}_{}{}^{j}v_{j}^{},\omega _i=v_i{}_{}{}^{\underset{¯}{j}}\omega _{\underset{¯}{j}}^{},$$ (C.3) we find that the spin connection components are $$\begin{array}{cccccc}\hfill \omega _{00i}& =& _i|M|,\hfill & \hfill \omega _{0ij}& =& \frac{1}{2}f_{ij},\hfill \\ & & & & & \\ \hfill \omega _{i0j}& =& \omega _{0ij},\hfill & \hfill \omega _{ijk}& =& |M|o_{ijk}2\delta _{i[j}_{k]}|M|,\hfill \end{array}$$ (C.4) where $`o_i^{jk}`$ is the 3-dimensional spin connection and $$_iv_i{}_{}{}^{\underset{¯}{j}}_{\underset{¯}{j}}^{},f_{ij}=v_i{}_{}{}^{\underset{¯}{k}}v_{j}^{}{}_{}{}^{\underset{¯}{l}}f_{\underset{¯}{k}\underset{¯}{l}}^{},f_{\underset{¯}{i}\underset{¯}{j}}2_{[\underset{¯}{i}}\omega _{\underset{¯}{j}]}.$$ (C.5) The components of the Riemann tensor are $$\begin{array}{ccc}\hfill R_{0i0j}& =& \frac{1}{2}_i_j|M|^2+_i|M|_j|M|\delta _{ij}(|M|)^2+\frac{1}{4}i|M|^6f_{ik}f_{jk},\hfill \\ & & \\ \hfill R_{0ijk}& =& \frac{1}{2}_i(|M|^4f_{jk})+\frac{1}{2}f_{i[j}_{k]}|M|^4\frac{1}{4}\delta _{i[j}f_{k]l}_l|M|^4,\hfill \\ & & \\ \hfill R_{ijkl}& =& |M|^2R_{ijkl}+\frac{1}{2}|M|^6(f_{ij}f_{kl}f_{k[i}f_{j]l})2\delta _{ij,kl}(|M|)^2+4|M|\delta _{[i}{}_{}{}^{[k}_{j]}^{}^{l]}|M|,\hfill \end{array}$$ (C.6) where all the objects in the right-hand sides of the equations are referred to the 3-dimensional spatial metric. The components of the Ricci tensor are $$\begin{array}{ccc}\hfill R_{00}& =& |M|^2^2\mathrm{log}|M|\frac{1}{4}|M|^6f^2,\hfill \\ & & \\ \hfill R_{0i}& =& \frac{1}{2}_j(|M|^4f_{ji}),\hfill \\ & & \\ \hfill R_{ij}& =& |M|^2\{R_{ij}+2_i\mathrm{log}|M|_j\mathrm{log}|M|\delta _{ij}^2\mathrm{log}|M|\frac{1}{2}|M|^4f_{ik}f_{jk}\},\hfill \end{array}$$ (C.7) and the Ricci scalar is $$R=|M|^2\{R\frac{1}{4}|M|^4f^22^2\mathrm{log}|M|+2(\mathrm{log}|M|)^2\},$$ (C.8) ## Appendix D Connection and curvature of a Brinkmann $`pp`$-wave metric We rewrite here for convenience the 4-dimensional form of these metrics: $$ds^2=2du(dv+Kdu+\omega )2e^{2U}dzdz^{},\omega =\omega _{\underset{¯}{z}}dz+\omega _{\underset{¯}{z}^{}}dz^{},$$ (D.1) where all the functions in the metric are independent of $`v`$. Using also light-cone coordinates in tangent space, a natural Vielbein basis is $$\begin{array}{cccccccccc}\hfill e^u& =& du\hfill & =& \widehat{l},\hfill & \hfill e_u& =& _{\underset{¯}{u}}K_{\underset{¯}{v}}\hfill & =& n^\mu _\mu ,\hfill \\ & & & & & & & & & \\ \hfill e^v& =& dv+Kdu+\omega \hfill & =& \widehat{n},\hfill & \hfill e_v& =& _{\underset{¯}{v}}\hfill & =& l^\mu _\mu ,\hfill \\ & & & & & & & & & \\ \hfill e^z& =& e^Udz\hfill & =& \widehat{m},\hfill & \hfill e_z& =& e^U(_{\underset{¯}{z}}\omega _{\underset{¯}{z}}_{\underset{¯}{v}})\hfill & =& m^\mu _\mu ,\hfill \\ & & & & & & & & & \\ \hfill e^z^{}& =& e^Udz^{}\hfill & =& \widehat{m}^{},\hfill & \hfill e_z^{}& =& e^U(_{\underset{¯}{z}^{}}\omega _{\underset{¯}{z}^{}}_{\underset{¯}{v}})\hfill & =& m^\mu _\mu .\hfill \end{array}$$ (D.2) The components of the spin connection are $$\begin{array}{cccccc}\hfill \omega _{uzu}& =& e^U(_{\underset{¯}{z}}K\dot{\omega }_{\underset{¯}{z}}),\hfill & \hfill \omega _{uzz^{}}& =& \frac{1}{2}e^{2U}f_{\underset{¯}{z}\underset{¯}{z}^{}}\dot{U},\hfill \\ & & & & & \\ \hfill \omega _{zz^{}u}& =& \frac{1}{2}e^{2U}f_{\underset{¯}{z}\underset{¯}{z}^{}}\dot{U},\hfill & \hfill \omega _{zzz^{}}& =& e^U_{\underset{¯}{z}}U,\hfill \end{array}$$ (D.3) where $`f_{\underset{¯}{z}\underset{¯}{z}^{}}=2_{[\underset{¯}{z}}\omega _{\underset{¯}{z}^{}]}`$ and a dot stands for partial derivation with respect to $`u`$. The components of the Ricci tensor are $$\begin{array}{ccc}\hfill R_{zz^{}}& =& 2e^{2U}_{\underset{¯}{z}}_{\underset{¯}{z}^{}}U,\hfill \\ & & \\ \hfill R_{zu}& =& \frac{1}{2}e^{3U}_{\underset{¯}{z}}f_{\underset{¯}{z}\underset{¯}{z}^{}}+e^U(_{\underset{¯}{z}}\dot{U}+\dot{U}_{\underset{¯}{z}}U),\hfill \\ & & \\ \hfill R_{uu}& =& 2e^{2U}_{\underset{¯}{z}}_{\underset{¯}{z}^{}}K+\frac{1}{2}(f_{\underset{¯}{z}\underset{¯}{z}^{}})^2+e^{2U}(_{\underset{¯}{z}}\dot{\omega }_{\underset{¯}{z}^{}}+_{\underset{¯}{z}^{}}\dot{\omega }_{\underset{¯}{z}})+2(\ddot{U}+\dot{U}\dot{U}),\hfill \end{array}$$ (D.4) and the Ricci scalar is just $$R=4e^{2U}_{\underset{¯}{z}}_{\underset{¯}{z}^{}}U.$$ (D.5)
warning/0506/cond-mat0506564.html
ar5iv
text
# Site and lattice resonances in metallic hole arrays ## Abstract A powerful analytical approach is followed to study light transmission through subwavelength holes drilled in thick perfect-conductor films, showing that full transmission (100%) is attainable in arrays of arbitrarily narrow holes as compared to the film thickness. The interplay between resonances localized in individual holes and lattice resonances originating in the array periodicity reveals new mechanisms of transmission enhancement and suppression. In particular, localized resonances obtained by filling the holes with high-index-of-refraction material are examined and experimentally observed through large enhancement in the transmission of individual holes. Light scattering from subwavelength holes drilled in metals has been the subject of long-standing interest Bethe (1944) motivated by phenomena such as extraordinary light transmission Chen (1971); R. C. McPhedran et al. (1980); Ebbesen et al. (1998) that challenges the severe $`(a/\lambda )^4`$ cut-off predicted by Bethe Bethe (1944) for the transmission cross section of single holes of small radius $`a`$ compared to the wavelength $`\lambda `$. In particular, 100% transmission was predicted to be attainable using lossless metals Chen (1971) and confirmed experimentally for quasi-perfect conductors in the microwave and THz domains R. C. McPhedran et al. (1980). Quite different from perfect-conductors, real metals are capable of sustaining surface plasmons that were readily recognized to mediate the interaction among arrayed holes at visible and near-infrared frequencies Ebbesen et al. (1998); Martín-Moreno et al. (2001); Barnes et al. (2004). Furthermore, the strong correlation of the transmission enhancement with the lattice periodicity in both of these metallic regimes has prompted rigorous descriptions of the effect in terms of dynamical diffraction Treacy (1999); Sarrazin et al. (2003); Lezec and Thio (2004), which connects directly to Wood’s anomalies Wood (1935); Sarrazin et al. (2003). Transmission resonances in individual holes Lezec et al. (2002); García de Abajo (2002); Falcone et al. (2004) offer an additional handle to achieve extraordinary effects. These resonances can be triggered by decorating a hole with a grating around it Lezec et al. (2002) (similar to some directional antennas designs Shafai and Kishk (1994)), by filling the hole with high-index-of-refraction material García de Abajo (2002), or by changing its shape to induce strong polarization Falcone et al. (2004); Klein Koerkamp et al. (2004). The combination of lattice resonances Collin and Eggimann (1961) in hole arrays and site resonances at specific hole positions can be anticipated to yield interesting properties in line with recent studies of light reflection on metal surfaces patterned with nanocavities that support localized modes Coyle et al. (2001); tra (a). In this Letter, we offer a systematics to study the phenomenology associated to light transmission through subwavelength hole arrays in perfect-conductor films, which permits us to establish the existence of full transmission resonances for arbitrarily small holes in thick metallic screens. Furthermore, individual-hole resonances are obtained by filling the holes with high-index-of-refraction materials. This gives rise to enhanced subwavelength-light transmission, which is demonstrated both theoretically and experimentally. Finally, the complex scenario that is presented when transmission resonances of individual holes are combined with resonances originating in the array periodicity is elucidated within our analysis. In his pioneering development, Bethe Bethe (1944) showed that the scattered far-field from a hole drilled in an infinitely-thin perfect-conductor screen can be assimilated to that of a magnetic dipole parallel to the screen plus an electric dipole perpendicular to it. Subsequent studies supplemented this result with higher-order multipole corrections Bouwkamp (1954), and eventually, with rigorous solutions for arbitrary hole radius and film thickness Roberts (1987); García de Abajo (2002). Small holes can be still represented by induced dipoles in thick screens, as illustrated in Fig. 1. This allows defining electric (E) and magnetic (M) polarizabilities both on the same side as the applied field ($`\alpha _\nu `$, with $`\nu =`$E,M) and on the opposite side ($`\alpha _\nu ^{}`$). Flux conservation under arbitrary illumination leads to an exact optical-theorem type of relationship between these polarizabilities: $`\mathrm{Im}\{g_\nu ^\pm \}=\mathrm{Im}\{{\displaystyle \frac{1}{\alpha _\nu \pm \alpha _\nu ^{}}}\}={\displaystyle \frac{2k^3}{3}},`$ (1) where $`k=2\pi /\lambda `$ is the momentum of light in free space. The remaining real parts of $`g_\nu ^\pm =(\alpha _\nu \pm \alpha _\nu ^{})^1`$ are obtained numerically Roberts (1987); García de Abajo (2002) and represented in Fig. 1b-c for empty holes. Note that $`\mathrm{Re}\{g_\nu ^+\}`$ diverges for zero thickness, in which case $`\mathrm{Re}\{\alpha _\nu \}=\mathrm{Re}\{\alpha _\nu ^{}\}`$ Bethe (1944), and that $`|\mathrm{Re}\{\alpha _\mathrm{M}^{}\}||\mathrm{Re}\{\alpha _\mathrm{E}^{}\}|`$ in the thick-film limit, dominated by transmission via the lowest-frequency TE guided mode, that does not create electric polarization. When a subwavelength hole is filled with dielectric material of sufficiently high permittivity $`ϵ`$, hole-cavity resonances can exist thanks to the reduction of the wavelength by a factor of $`\sqrt{ϵ}`$. These resonances give rise to enhanced transmission García de Abajo (2002), as shown in Fig. 2a by rigorous numerical solution of Maxwell’s equations (curves) Roberts (1987); García de Abajo (2002). We present experimental evidence of this hole-resonance behavior in Fig. 2b, which compares the transmission of microwaves through subwavelength holes filled with teflon and air. A 5-fold enhancement in the transmission is observed. The width of these resonances is dictated by coupling of the cavity modes to the continuum of light states outside the film. The resonances are of Fabry-Perot origin tra (b), but the transmission line shapes are actually determined from the noted coupling to the two continua outside the film, as described by Fano tra (b) (the vanishing of the transmission when $`g_\mathrm{M}^+=g_\mathrm{M}^{}`$, i.e., $`\alpha _M^{}=0`$, is a signature of a Fano resonance; see below). The coupling strength drops rapidly for large $`ϵ`$ due in part to small transmission through the dielectric-air interface, as predicted by Fresnel’s equations. The larger $`ϵ`$, the narrower the resonance, and the higher the transmission maxima. Incidentally, the normalized transmission cross-section obtained from our effective dipole model ($`16k^4|\alpha _\mathrm{M}^{}|^2/3a^2`$, symbols in Fig. 2a for $`ϵ=50`$) compares remarkably well with the exact result (curve) for small $`a/\lambda `$. Periodic arrays of sufficiently small and spaced holes can also be described by perpendicular electric dipoles $`p`$ and $`p^{}`$ and parallel magnetic dipoles $`m`$ and $`m^{}`$, where primed (unprimed) quantities are defined on the entry (exit) side of the film as determined by the incoming light. This is an extension of previous considerations for thin screens relying on Babinet’s principle Collin and Eggimann (1961); tra (b). We consider first a unit-electric-field p-polarized plane wave incident on a hole array with parallel momentum $`𝐤_{}`$ along the $`x`$ axis, so that the external (incident plus reflected) field in the absence of the holes has parallel magnetic field $`H_y^{\mathrm{ext}}=2`$ along the $`y`$ direction and perpendicular electric field $`E_z^{\mathrm{ext}}=2k_{}/k`$ along $`z`$. One can write the following set of multiple-scattering equations for the self-consistent dipoles, that respond both to the external field and to the field scattered by the other holes: $`p`$ $`=`$ $`\alpha _\mathrm{E}(E_z^{\mathrm{ext}}+G_zpHm)+\alpha _\mathrm{E}^{}(G_zp^{}Hm^{})`$ $`p^{}`$ $`=`$ $`\alpha _\mathrm{E}^{}(E_z^{\mathrm{ext}}+G_zpHm)+\alpha _\mathrm{E}(G_zp^{}Hm^{})`$ $`m`$ $`=`$ $`\alpha _\mathrm{M}(H_y^{\mathrm{ext}}+G_ymHp)+\alpha _\mathrm{M}^{}(G_ym^{}Hp^{})`$ $`m^{}`$ $`=`$ $`\alpha _\mathrm{M}^{}(H_y^{\mathrm{ext}}+G_ymHp)+\alpha _\mathrm{M}(G_ym^{}+Hp^{}),`$ where $`G_j`$ and $`H`$ describe the induced fields produced at a given hole by the other holes Collin and Eggimann (1961). Noticing that the dipoles depend on hole positions $`𝐑=(x,y)`$ only via phase factors $`\mathrm{exp}(\mathrm{i}k_{}x)`$, one finds $`G_j`$ $`=`$ $`{\displaystyle \underset{𝐑0}{}}\mathrm{e}^{\mathrm{i}k_{}x}(k^2+_j_j){\displaystyle \frac{\mathrm{e}^{\mathrm{i}kR}}{R}}`$ (2) $`H`$ $`=`$ $`\mathrm{i}k{\displaystyle \underset{𝐑0}{}}\mathrm{e}^{\mathrm{i}k_{}x}_x{\displaystyle \frac{\mathrm{e}^{\mathrm{i}kR}}{R}}.`$ The solution of the above equations can be written $`p\pm p^{}`$ $`=`$ $`2[(g_\mathrm{M}^\pm G_y)k_{}/k+H]/\mathrm{\Delta }_\pm `$ $`m\pm m^{}`$ $`=`$ $`2[(g_\mathrm{E}^\pm G_z)+Hk_{}/k]/\mathrm{\Delta }_\pm `$ with $`\mathrm{\Delta }_\pm =(g_\mathrm{E}^\pm G_z)(g_\mathrm{M}^\pm G_y)H^2,`$ (3) from where the zeroth-order transmission of the holey film can be evaluated as obtained from the far field set up by an infinite 2D array of dipoles: $`T=|{\displaystyle \frac{2\pi k^2}{Ak_z}}(m^{}p^{}k_{}/k)|^2.`$ Here, $`A`$ is the lattice unit-cell area and $`k_z=\sqrt{k^2k_{}^2}`$. Similarly, the transmittance of s-polarized light reduces to $`T=|2\pi km^{}/A|^2`$, with magnetic dipoles parallel to $`𝐤_{}`$ and no electric dipoles whatsoever ($`E_z^{\mathrm{ext}}=0`$). More precisely, $`m\pm m^{}`$ $`=`$ $`{\displaystyle \frac{2k_z/k}{g_\mathrm{M}^\pm G_x}},`$ from where one obtains $`T`$ $`=({\displaystyle \frac{2\pi k_z}{A}})^2|{\displaystyle \frac{1}{g_\mathrm{M}^+G_x}}{\displaystyle \frac{1}{g_\mathrm{M}^{}G_x}}|^2`$ $`=`$ $`|{\displaystyle \frac{1}{1+\frac{\mathrm{i}A}{2\pi k_z}\mathrm{Re}\{g_\mathrm{M}^+G_x\}}}{\displaystyle \frac{1}{1+\frac{\mathrm{i}A}{2\pi k_z}\mathrm{Re}\{g_\mathrm{M}^{}G_x\}}}|^2.`$ The last identity in Eq. (Site and lattice resonances in metallic hole arrays) is derived from Eq. (1) and from the exact relation $`\mathrm{Im}\{G_x\}=2\pi k_z/A2k^3/3`$ for propagating light ($`k_{}<k`$). The performance of the hole array is dominated by divergences in the lattice sums when the diffraction orders $`(m,n)`$ go grazing. More precisely, for a square lattice of spacing $`d`$ and for $`𝐤_{}`$ along $`x`$, the sums $`G_j`$ go to $`+\mathrm{}`$ as $`G_j{\displaystyle \frac{1}{\sqrt{(k_{}+2\pi m/d)^2+(2\pi n/d)^2k^2}}}.`$ (5) In particular, $`G_y`$ and $`G_z`$ (p polarization) diverge on the lowest-frequency side of all grazing diffraction orders, as illustrated in Fig. 3, whereas $`G_x`$ (s polarization) diverges only for $`n0`$ (non-straight curves). This entails different peak structure patterns for s- and p-polarized light (see Fig. 4). Interestingly, Eq. (Site and lattice resonances in metallic hole arrays) predicts 100% transmission whenever the condition $`1+({\displaystyle \frac{A}{2\pi k_z}})^2\mathrm{Re}\{g_\mathrm{M}^+G_x\}\mathrm{Re}\{g_\mathrm{M}^{}G_x\}=0`$ (6) is fulfilled. Eq. (6) is a second-order algebraic equation in $`\mathrm{Re}\{G_x\}`$ that admits positive real solutions (one or two depending on the signs of $`\mathrm{Re}\{g_\mathrm{M}^\pm \}`$) when $`{\displaystyle \frac{A}{4\pi k_z}}|g_M^+g_M^{}|>1,`$ (7) a condition that can be easily satisfied near $`n0`$ grazing diffraction orders, where $`G_x`$ can be chosen arbitrarily large within a narrow range of wavelengths \[see Eq. (5)\]. It should be noted that the difference $`g_M^+g_M^{}`$ falls off rapidly to zero when the film thickness $`t`$ is made much larger than the hole radius for empty holes (see Fig. 1b). However, for fixed $`t/a`$ ratio and angle of incidence, the left hand side of (7) reduces to a positive real constant times $`\lambda A/a^3`$, leading to the conclusion that 100% transmission is possible regardless how small the holes are as compared to the film thickness, provided the separation between holes (or equivalently $`A`$) is made sufficiently large. The interaction between site and lattice resonances is explored in Fig. 4 through the transmittance of square lattices of holes filled with materials of different permittivity for various values of the $`t/a`$ ratio, both for p-polarized and s-polarized incident light. The dominant features of these plots can be classified as follows: (i) Full transmission close to lattice resonances as those of Fig. 3. In particular under the conditions of Fig. 4d, one can neglect the first term inside the squared modulus of Eq. (Site and lattice resonances in metallic hole arrays) ($`\mathrm{Re}\{g_\mathrm{M}^+\}>>\mathrm{Re}\{g_\mathrm{M}^{}\}`$, see Fig. 1b), so that 100% transmittance maxima come about near $`n0`$ grazing diffraction orders (see Fig. 3) for which one can have $`\mathrm{Re}\{g_\mathrm{M}^{}\}\mathrm{Re}\{G_x\}`$. (ii) Full transmission close to dispersionless site resonances. This is illustrated graphically in the left part of Fig. 5, which shows that features A and C of Fig. 4j correspond to 100% transmission at wavelengths where the condition noted in (i) is fulfilled. The resonant individual-hole polarizabilities represented in the lower part of Fig. 5 display a typical Lorentzian shape in coincidence with a transmission maximum for isolated holes (see $`ϵ=100`$ curve in Fig. 2a), from which the full transmission maximum in the lattice is blue-shifted due to inter-hole interaction described by $`G_x`$. The density of site resonances increases dramatically both with film thickness (see Fig. 4l) and with $`ϵ`$ (Fabry-Perot-like behavior). (iii) Dispersionless regions of vanishing transmission. Eq. (Site and lattice resonances in metallic hole arrays) predicts $`k_{}`$-independent vanishing transmission when $`g_\mathrm{M}^+=g_\mathrm{M}^{}`$, which is a property of single holes. This is the case of feature B in Fig. 4j, as illustrated geometrically in the right part of Fig. 5. (iv) Strong mixing of site and lattice resonances. Avoided level crossings are particularly evident in Figs. 4c,g near $`k_{}=\pi /d`$. Non-avoided crossings are also observed, as well as splitting of full transmission maxima. (v) Film-bound states. For incident evanescent light with $`k<k_{}<2\pi /dk`$, the lattice sums satisfy $`\mathrm{Im}\{G_j\}=2k^3/3`$ and $`\mathrm{Im}\{H\}=0`$. This implies that $`\mathrm{\Delta }_\pm `$ \[Eq. (3)\] is real and can vanish for specific combinations of $`k`$ and $`k_{}`$, leading to simultaneous infinite transmittance and reflectance (evanescent waves do not propagate energy) in what constitute film-bound resonances, as recently predicted for related metal structures Pendry et al. (2004). In summary, a simple and powerful formalism has been used to analyze transmission through hole arrays leading to surprising results such as 100% transmission for thick perfect-conductor films perforated by arbitrarily small holes. Both theoretical and experimental evidence of single-hole resonances obtained by filling the holes with large-index-of-refraction material have been presented, resulting in enhanced transmission through isolated holes. Finally, filled-hole arrays have been shown to exhibit a colorful phenomenology, including new types of suppressed transmission and a complicated interplay between hole-site resonances and lattice resonances that is explained within the present approach.
warning/0506/cond-mat0506574.html
ar5iv
text
# Semiclassical ordering in the large-𝑁 pyrochlore antiferromagnet ## Abstract We study the semiclassical limit of the $`Sp(N)`$ generalization of the pyrochlore lattice Heisenberg antiferromagnet by expanding about the $`N\mathrm{}`$ saddlepoint in powers of a generalized inverse spin. To leading order, we write down an effective Hamiltonian as a series in loops on the lattice. Using this as a formula for calculating the energy of any classical ground state, we perform Monte Carlo simulations and find a unique collinear ground state. This state is not a ground state of linear spin-wave theory, and can therefore not be a physical ($`N=1`$) semiclassical ground state. Geometrically frustrated antiferromagnets Ramirez (1994); Diep (2005) have attracted interest because their large classical ground state degeneracy can allow a rich variety of correlated states, including (at $`T=0`$) quantum spin liquids or complex ordered states. The simplest examples are nearest-neighbor, exchange-coupled antiferromagnets in which spins from triangles or tetrahedra that share corners Henley (2001a): the kagomé, “checkerboard”, and SCGO ($`\mathrm{SrCr}_{9\mathrm{p}}\mathrm{Ga}_{129\mathrm{p}}\mathrm{O}_{19}`$) lattices Ramirez (1994) in two dimensions, plus the garnet and pyrochlore lattices in three dimensions: the pyrochlore, in particular, consists of tetrahedra whose centers form a diamond lattice. The Hamiltonian is $`H=J_{ij}𝐒_i𝐒_j`$, where $`J_{ij}=1`$ for nearest neighbors $`ij`$. In fact, additional terms – dipole interactions and anisotropies (as in $`\mathrm{Gd}_2\mathrm{Ti}_2\mathrm{O}_7`$), magnetoelastic couplings (as in $`\mathrm{ZnV}_2\mathrm{O}_4`$ and $`\mathrm{ZnCr}_2\mathrm{O}_4`$)– decide the order in most real materials Ramirez (1994); Diep (2005). Still, the case with pure Heisenberg exchange is worth understanding since (i) most simulations are done for this case; (ii) the more realistic systems emerge from it by the addition of perturbations; (iii) this has motivated experimentalists to search for model systems in which the aforementioned perturbations are small; (iv) quantum effects can be studied without being overshadowed by classical effects. What is the ground state for large spin length $`S`$? In unfrustrated antiferromagnets, it is just the classical ground state dressed with zero-point fluctuations of harmonic spin waves, and in frustrated cases the spin-wave zero-point energy *may* lift the degeneracy of classical ground states Shender (1982). In the pyrochlore case, though, a large degeneracy remains Henley (2005b); its resolution by higher-order (anharmonic) terms in the semiclassical ($`1/S`$) expansion requires arduous approximations Hizi and Henley (2005a). An established alternative to the spin-wave approach is to generalize the Heisenberg spins \[with SU($`2`$) $``$ Sp($`1`$) symmetry\] to Sp($`N`$) symmetry Read and Sachdev (1991): here $`N`$ is the number of flavors of Schwinger bosons whose bilinear form represents a *generalized spin* Read and Sachdev (1991), with length $`\kappa =2S`$. The resulting mean-field theory (valid in the $`N\mathrm{}`$ limit) is popular as an analytic approach to the $`S=1/2`$ limit, since the *small*-$`\kappa `$ limit captures various disordered and exotic ground states Read and Sachdev (1991); Bernier et al. (2005). The large-$`N`$ mean-field theory is also useful at *large*-$`\kappa `$ for this paper’s problem, since it gives a simple analytical prescription for ground state selection: unlike the spin-wave expansion, here all degeneracies are (typically) broken at the lowest order \[$`𝒪(1/\kappa )]`$ quantum correction Sachdev (1992); Read and Sachdev (1991). However, on highly frustrated lattices this approach has the complication of a macroscopic (exponential) number of degenerate saddle-points not related by symmetry, so it is unknown a priori which of these should be expanded around; this was handled till now by limiting the investigation to ordering patterns of high symmetry and small magnetic cells, or by enumerating all saddle-points in a small finite system Sachdev (1992); Bernier et al. (2004). In this letter, we develop an *effective Hamiltonian* Henley (2001a) approach to this question. The pertinent saddle-points are labeled by arrangements of valence bond variables, and we obtain a simple formula for the large-$`N`$ mean-field energy of *any* classical ground state, as a function of these variables. The effective Hamiltonian is constructed as an analytical real-space expansion of *loops* made of valence bonds. This allows us to systematically search for a collinear pyrochlore ground state, using Monte Carlo annealing, on quite large system sizes. However, we also find that the pyrochlore ground state does *not* agree with even the lowest-order term in the spin-wave expansion, and therefore cannot give the right answer for the physical ($`N=1`$) ground state, in the large-$`S`$ limit, demonstrating a limitation of the large-$`N`$ approach for this case. *Large $`N`$ mean field theory.*—We begin by discussing the mean-field Hamiltonian derived from the Sp($`N`$) generalization of $`H`$. For the $`N=1`$ case we can write the spin interaction in terms of Schwinger boson operators as $`\stackrel{}{S}_i\stackrel{}{S}_j=b_{i\sigma }^{}b_{i\sigma ^{}}b_{j\sigma ^{}}^{}b_{j\sigma }`$, where a sum over repeated indices $`\sigma `$ and $`\sigma ^{}`$ (that take values $`,`$) is implied. The Hilbert space of the spin model is obtained by constraining the number of bosons on each site $`b_{i\sigma }^{}b_{i\sigma }=2S`$. We can rewrite the interaction in terms of valence bonds created by the operator $`ϵ_{\sigma \sigma ^{}}b_{i\sigma }^{}b_{j\sigma ^{}}^{}`$, where $`ϵ_{}=ϵ_{}=1`$. An arbitrary singlet state can be written in terms of some arrangement of these bonds with at most $`2S`$ bonds emanating from any lattice site. Generalizing these bond operators to $`N`$-flavors allows us to put a large number of bonds on a link. Since the Hamiltonian acting on a state changes at most two bonds per link, the relative change in the number of bonds goes like $`1/NS`$. In the large-$`N`$ limit, their fluctuations are quenched. Therefore, we factorize the interaction in terms of valence bonds $`Q_{ij}=ϵ_{\sigma \sigma ^{}}b_{i\sigma ,m}b_{j\sigma ^{},m}/N`$, where the *flavor* index $`m=1,2\mathrm{},N`$. We treat $`Q_{ij}`$ as classical quantities, to obtain the mean-field Hamiltonian $`H_{\mathrm{MF}}`$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle \underset{ij}{}}[N|Q_{ij}|^2+(ϵ_{\sigma \sigma ^{}}b_{i\sigma ,m}^{}b_{j\sigma ^{},m}^{}Q_{ij}+H.c.)]`$ (1) $`+`$ $`{\displaystyle \underset{i}{}}\lambda _i\left(b_{i\sigma ,m}^{}b_{i\sigma ,m}N\kappa \right)`$ Here Lagrange multipliers $`\lambda _i`$ have been introduced to enforce the constraint on boson number $`N\kappa `$ at every site $`i`$, defining the *generalized spin length* $`\kappa =2S`$. In what follows, we shall take $`\lambda _i`$ to be spatially uniform $`\lambda _i=\lambda `$. We are interested in large enough values of $`\kappa `$ to condense a flavor mode of the itinerant bosons, $`b_{i\sigma ,m}=\sqrt{N}\delta _{1,m}x_{i\sigma }`$, for long-range order to develop. The mean-field ground state energy (per flavor) is obtained by diagonalizing (1) by a canonical Bogoliubov transformation: $`{\displaystyle \frac{E_{\mathrm{MF}}}{N}}`$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle \underset{ij}{}}[|Q_{ij}|^2+(ϵ_{\sigma \sigma ^{}}x_{i\sigma }x_{j\sigma ^{}}Q_{ij}^{}+c.c.)]`$ (2a) $`+`$ $`{\displaystyle \underset{i}{}}\lambda \left(|x_{i\sigma }|^2\kappa \right)`$ $`+`$ $`{\displaystyle \frac{1}{2}}[\mathrm{Tr}\sqrt{\lambda ^2𝟙𝐐^{}𝐐}N_s\lambda ]`$ (2b) Here $`N_s`$ is the number of lattice sites, and (2b) is the zero-point energy contribution of the bosons. The exact mean-field ground state is obtained by a constrained minimization of the above expression. It can be systematically approached as an expansion in powers of $`1/\kappa `$. The leading contribution to the energy (of order $`\kappa ^2`$) comes from terms in (2a), whose minimization simply relates the valence bonds to the condensate configuration in the classical ground state(s) of the Heisenberg Hamiltonian $`H`$ with spin size $`\kappa /2`$. We will denote this configuration of bond variables with a superscript $`c`$: $`\{Q_{ij}^c\}`$. The quantum correction (of order $`\kappa `$) is provided by terms in (2b) for these bond configurations. The ground states of the classical Hamiltonian (2a) consist of all spin configurations in which the spin vectors sum to zero in every tetrahedron. On general grounds we expect quantum corrections to select *collinear* ground states from the classical manifold Shender (1982); Henley (2001a); Hizi and Henley (2005b). We therefore restrict our attention to such states, in which each spin can be denoted by an Ising variable $`\eta _i\{\pm 1\}`$. Collinearity implies that, up to an arbitrary gauge transformation, $`Q_{ij}^c=\kappa (\eta _i\eta _j)/2`$ and thus the bond variables are $`\pm \kappa `$ for every satisfied, antiferromagnetic (AFM) bond, and zero otherwise. Also, $`\lambda ^c=4\kappa `$ for all pyrochlore lattice classical ground states. *Loop expansion and effective Hamiltonian.*—Next,we recast the first quantum correction to the mean field energy, Eq. (2b), for a given collinear classical ground state, into an effective Hamiltonian form where only some of the degrees of freedom remain Henley (2001a). Eq. (2b) can formally be Taylor-expanded $$\frac{E_q}{N}=\frac{1}{2}\underset{m=1}{\overset{\mathrm{}}{}}\frac{(2m+1)!!}{2^m\lambda ^{2m1}m!}\mathrm{Tr}\left(𝐐^{}𝐐\right)^m$$ (3) Since $`|Q_{ij}/\kappa |=1`$ for AFM bonds, and zero otherwise, $`\mathrm{Tr}(𝐐^{}𝐐/\kappa ^2)^m`$ is equal to the number of closed paths of length $`2m`$, composed of AFM bonds. All terms in Eq. (3) depend solely on the structure of the network formed by AFM bonds. Note that since this network is bipartite, each nonzero element of $`𝐐^{}𝐐`$ is $`\kappa ^2`$. In any collinear classical ground state, each tetrahedron has two up spins and two down spins, and four AFM bonds forming a closed loop (see Figs. 1a-d). This means that the local connectivity of the AFM network is identical for all states, and many closed paths only contribute state-independent terms to Eq. (3). For example, $`\mathrm{Tr}𝐐^{}𝐐=4N_s\kappa ^2`$, for any classical ground state since the only paths of length $`2`$ involve going to and fro on the same bond, and each site has four neighbors which have the opposite spin (see Fig. 1a). Similarly $`\mathrm{Tr}(𝐐^{}𝐐)^2=(16+12+4)N_s\kappa ^4`$, where the three terms correspond to the paths shown in Figs. 1b, 1c, 1d, respectively. All paths that do not involve loops, (e.g. those in Figs. 1a, 1b, 1c) can be viewed as paths on a Bethe lattice of coordination $`4`$, and would contribute a constant term to the energy for all collinear classical ground states. The same is true for paths involving only trivial loops, in addition to the Bethe lattice path, as in Fig. 1d. Here, a “trivial” loop is the loop of length $`4`$ that exists within any tetrahedron. The lowest order terms in expansion (3) that contribute a state-dependant term in the effective Hamiltonian are for $`2m=6`$, since the shortest non-trivial loops are hexagons. This leads us to parameterize the effective Hamiltonian in terms of the various non-trivial AFM loops. $$\frac{E_q^{\mathrm{eff}}}{N(\kappa /2)}=K_0+K_6𝒫_6+K_8𝒫_8+K_{10}𝒫_{10}+\mathrm{},$$ (4) where $`\{K_{2l}\}`$ are numerical coefficients, and $`𝒫_{2l}`$ is the number of non-trivial AFM loops of length $`2l`$, per site. To evaluate the coefficients $`\{K_{2l}\}`$, we need to calculate two types of terms: (i) The number $`F(2m)`$ of closed paths of total length $`2m`$ on a decorated (with trivial $`4`$-loops) coordination-$`4`$ Bethe lattice (Fig. 1e). (ii) The number $`G(2l,2m)`$ of closed paths of length $`2(m+l)`$, involving a particular loop of length $`2l`$ with decorated Bethe lattice paths emanating from each site along the loop (Fig. 1f). Calculating these terms is a matter of tedious but tractable combinatorics. We find that the functions $`F(2m)`$, $`G(2l,2m)`$ decay rapidly with $`m`$, allowing us to sum them in order to evaluate the coefficients to any accuracy in Eq. (4), using $$K_0=\underset{m=0}{\overset{\mathrm{}}{}}F(2m),K_{2l}=\underset{m=0}{\overset{\mathrm{}}{}}G(2l,2m).$$ (5) We show the first five coefficients in Tab. 1. Thus we have obtained an effective Hamiltonian that is parameterized solely by the number of AFM loops of various sizes. Note that the coefficients decay rapidly $`K_{2l+2}/K_{2l}1/10`$, which leads us to expect short loops to be the dominant terms in the expansion. This allows us, in principle, to calculate the energy, to any accuracy, for any member of an *infinite* ensemble of classical ground states. This represents a significant improvement over previous calculations that were always limited to small system sizes Sachdev (1992); Bernier et al. (2004). Although we derived the effective Hamiltonian for collinear states, it turns out that, in fact, the classical tetrahedron zero sum rule implies that Eq. (4), with the coefficients in Tab. 1, is valid for *any non-collinear* classical ground state, as well, with the generalized loop variables expressed as sums over non-trivial loops $$𝒫_{2l}=\frac{1}{\kappa ^{2l}}\underset{(i_1\mathrm{}i_{2l})}{}\mathrm{Re}(Q_{i_1i_2}^{}Q_{i_2i_3}\mathrm{}Q_{i_{2l1}i_{2l}}^{}Q_{i_{2l}i_1}).$$ (6) Unlike the collinear case, where the elements of $`𝐐^{}𝐐`$ could only take the values $`0`$ or $`\kappa ^2`$, and thus each loop would contribute $`0`$ or $`1`$ to the sum (6), in the general case, the matrix elements of $`𝐐^{}𝐐`$ are complex. *Numerical results.*—To verify the validity of the effective Hamiltonian (4), we calculated the energy for a large number of collinear classical ground states, as well as linear spin-wave ground states, obtained by a random flipping algorithm described elsewhere Hizi and Henley (2005b). We find that the energies are remarkably well described by $`E_q^{\mathrm{eff}}`$, even when we cut the expansion (4) off at $`2l=8`$, as shown in Fig. 2. We used the coefficient values of Tab. 1, but had to adjust the constant term $`K_0`$ *separately* for each choice of cutoff, in order to get a good fit footnote . In practice, this means that the effective Hamiltonian (4) is extremely useful for *comparing* energies of various states, even with a small cutoff, but requires many terms in order to accurately determine the energy. An independent $`5`$-parameter numerical fit, to Eq. (4), up to $`2l=12`$, gives the values shown in the right-hand column of Tab. 1. Now that we have an approximate formula for $`E_q`$, for any collinear classical ground state, we can systematically search these states, with large magnetic unit cells, to find a ground state. We conducted Monte Carlo simulations using a Metropolis loop flipping algorithm and the effective energy of Eq. (4), for various orthorhombic unit cells of sizes ranging from $`128`$ to $`3456`$ sites, with periodic boundary conditions. We find a minimum energy of $`E_q/(N\kappa /2)=0.60077N_s`$ for a family of *nearly* degenerate states. They are composed of layers, that can each be in one of four arrangements, resulting in $`e^{cL}`$ states, where $`L`$ is the system size, and $`c`$ is a constant. Each of these states has $`𝒫_6=N_s/3`$, which is the maximum value that we find (but is *not* unique to these states), and $`𝒫_8=23N_s/6`$. Upon closer investigation, however, we find that a *unique ground state* (depicted in Fig. 3) is selected. The energy difference to nearby states is of order $`10^7N_s`$, corresponding to the $`2l=16`$ term. *Discussion.*—It was noted by one of us Henley (2005b) (see also Tchernyshyov (2004)) that, in the pyrochlore, the degeneracy of ground states of the spin-wave quantum Hamiltonian, at the lowest order in $`1/S`$, is associated with a *gauge-like* symmetry. This symmetry characterizes the degenerate sub-manifold of collinear spin ground states by the condition $$\underset{i\mathrm{}}{}\eta _i=1,$$ (7) for all non-trivial hexagons. Since the spin-wave theory is expected to be exact in the limit of infinite $`S`$, the physical ground state must satisfy Eq. (7). The state depicted in Fig. 3, however, *does not*. Looking at the inset in Fig. 2, we find that states with negative hexagon products tend to have lower large-$`N`$ energy than other states, since they tend to have more AFM loops, but this is not a strict rule. Thus it would seem that the $`N\mathrm{}`$, large-$`\kappa `$ ground state cannot be the physical ($`N=1`$) large-$`S`$ ground state. Nevertheless, if we restrict the large-$`N`$ calculation to harmonic spin-wave ground states only, we find that the ordering of energies for various states is similar to preliminary anharmonic spin-wave results Hizi and Henley (2005a), and does predict the same ground state. As shown in Fig. 4, in both cases, the lowest energy among harmonic spin-wave ground state belongs to a state with the most AFM hexagons. The effective Hamiltonian approach that we have outlined here can easily be applied to other lattices. In the checkerboard lattice, the energy is lowest for states that have the most AFM (square) non-trivial plaquettes. Thus, the non-degenerate ground state is clearly the ($`\pi `$,$`\pi `$) state in which *all* plaquettes are AFM Bernier et al. (2004). In the kagomé case, all classical ground states are non-collinear. However, if we limit ourselves to coplanar arrangements, we find that $`Q_{ij}`$ has the same absolute value for *all* of the lattice bonds, but the signs differ depending on the *chirality* of the triangle to which the bond $`(i,j)`$ belongs. Therefore, the effective Hamiltonian (4), with the generalized variables (6), prefers classical ground states with negative product of triangle chiralities around all hexagons. One can thus conclude that the ground state is the $`\sqrt{3}\times \sqrt{3}`$ state, as large-$`N`$ calculations have indeed found Sachdev (1992). Let us also remark that our method can be generalized to long-range Heisenberg interactions which are relevant in the context of real materials like $`\mathrm{Tb}_2\mathrm{Ti}_2\mathrm{O}_7`$ Ramirez (1994); Diep (2005). Finally, it has been suggested that the disordered (*small*-$`\kappa `$) limit of the large-$`N`$ approximation for the pyrochlore lattice also has a massive multiplicity of saddle-points Tchernyshyov et al. (2004); an effective Hamiltonian similar to this paper’s could organize the handling of this family. ###### Acknowledgements. UH and CLH acknowledge support from NSF grant DMR-0240953, and PS acknowledges support from the Packard foundation and the U.S. Dept. of Energy, under Contract No. W-31-109-ENG-38.
warning/0506/astro-ph0506624.html
ar5iv
text
# Exotica in rotating compact stars ## 1 Introduction Neutron stars are ideal astrophysical laboratories to study properties of cold and dense matter . The investigation of dense and cold matter in neutron star core has entered into a new phase. Observatories such as Chandra X-ray observatory, X-ray Multi Mirror (XMM)-Newton, Rossi X-ray Timing Explorer (RXTE) are pouring in many new and interesting data on compact stars. It is now possible to estimate mass and radius of a single neutron star from observations. These informations may shed light on the composition and equation of state (EoS) of matter at low temperature and high baryon chemical potential. A lot of interest has been generated in thermonuclear x-ray bursts from the low mass x-ray binary EXO 0748-676. Already Cottam et al. estimated the gravitational redshift $`z=`$ 0.35 of three absorption lines in x-ray bursts from EXO 0748-676 with XMM-Newton. They also reported the narrow widths of those lines. Mass to radius ratio is immediately known from the measurement of $`z=(12GM/c^2R)^{1/2}1`$, where M and R are neutron star mass and radius. On the other hand, the widths of absorption lines provide information about the surface rotational velocity ($`v_{rot}`$) which is related to radius and spin frequency $`v_{rot}\nu _{spin}R`$. Villarreal and Strohmayer reported the neutron star spin frequency ($`\nu _{spin}`$) of 45 Hz in the x-ray burst oscillations of EXO 0748-676 with RXTE. These observations lead to a radius of 9.5-15 km and mass 1.5-2.3M for EXO 0748-676 . For the first time, it has been possible to determine mass and radius of a single neutron star. Recently discovered double pulsar system PSR J0737-3039 provides another opportunity for the determination of mass and radius of same neutron star. Two pulsars in this double pulsar system are denoted by pulsar A and pulsar B. Pulsar A has a spin period of $``$ 23 ms and mass 1.34M whereas those of pulsar B are 2.8 s and 1.25M. Double neutron star binaries provide unique laboratories for testing relativistic gravity. Soon it will be possible to measure spin-orbit coupling of pulsar A from pulsar timing data . The moment of inertia for pulsar A could be determined from the measurement of spin-orbit coupling. Moment of inertia is dimensionally proportional to mass times radius squared. Already, the mass of pulsar A is known. Therefore, the measurement of moment of inertia will provide an estimate of radius of the compact object and put important constrain on the properties of neutron star matter. Now the observed mass-radius (M-R) relationship can be directly compared with that of theoretical calculation. Consequently, the composition and EoS of dense matter in neutron stars may be probed. Here we discuss exotic forms of matter and their influence on various properties of static and rotating compact stars. In section 2, we describe the models of static and uniformly rotating compact stars and EoS adopted for this calculation. Results are discussed in section 3. And section 4 is devoted to summary and conclusions. ## 2 Model Models of static and rotating compact stars are constructed by solving Einstein’s field equations. The equilibrium structures of non-rotating stars are obtained from Tolman-Oppenheimer-Volkoff (TOV) equations. For rotating neutron stars, we consider stationary and axisymmetric equilibrium configurations. In this case, the metric is taken in the form $$ds^2=e^{\gamma +\rho }dt^2+e^{2\alpha }(dr^2+r^2d\theta ^2)+e^{\gamma \rho }r^2sin^2\theta (d\varphi \omega dt)^2,$$ (1) where $`\omega `$ is the angular velocity of the local inertial frame. The metric functions in the line element depend on the radial coordinate $`r`$ and the polar angle $`\theta `$. We model the neutron star matter as a perfect fluid and the stress-energy tensor is given by $$T^{\mu \nu }=(\epsilon +P)u^\mu u^\nu Pg^{\mu \nu }$$ (2) where $`\epsilon `$ and $`P`$ are the total energy density and pressure of the matter respectively and $`u^a`$ is the fluid’s four velocity. The moment of inertia of the star is defined by $`I=J/\mathrm{\Omega }`$; angular momentum ($`J`$) is, $$J=d^3x\sqrt{(}g)(ϵ+P)u^t(g_{\varphi \varphi }u^\varphi +g_{\varphi t}u^t)$$ (3) where, $`u^\varphi `$ and $`u^t`$ are components of fluid four velocity and $`\mathrm{\Omega }`$ is angular velocity of the star. For uniformly rotating neutron stars, we adopt a numerical code based on Komatsu-Eriguchi-Hachisu method. Composition and structure of neutron stars depend on the nature of strong interaction. Matter density could exceed by a few times normal nuclear matter density in a neutron star interior. Consequently, the baryon and lepton chemical potential increase rapidly with baryon density. This might lead to the formation of exotic matter with a large strangeness fraction such as hyperon matter, Bose-Einstein condensate of $`K^{}`$ mesons and quark matter. Here we discuss equations of state with various compositions of dense matter in neutron star cores. We construct hadronic EoS using relativistic field theoretical models . Here baryon-baryon interaction is mediated by exchange of mesons. The hadronic phase is composed of all members of the baryon octet, electrons and muons. The hadronic phase is to satisfy beta-equilibrium and charge neutrality conditions. In compact stars interior, the generalised $`\beta `$-decay processes may be written in the form $`B_1B_2+l+\overline{\nu }_l`$ and $`B_2+lB_1+\nu _l`$ where $`B_1`$ and $`B_2`$ are baryons and l represents leptons. The generic equation for chemical equilibrium condition is $$\mu _i=b_i\mu _nq_i\mu _e,$$ (4) where $`\mu _n`$, $`\mu _e`$ and $`\mu _i`$ are respectively the chemical potentials of neutrons, electrons and i-th baryon and $`b_i`$ and $`q_i`$ are baryon number and electric charge of i-th baryon respectively. The charge neutrality in hadronic phase is given by $`Q^h=_Bq_Bn_B^hn_en_\mu =0`$, where $`n_B^h`$ is the number density of baryon B in pure hadronic phase and $`n_e`$ and $`n_\mu `$ are number densities of electrons and muons respectively. Next we consider Bose-Einstein condensate of $`K^{}`$ mesons. Experimental results from heavy ion collisions suggest that the in-medium kaon-nucleon interaction is repulsive whereas it is attractive for antikaon-nucleon . Also, informations about antikaon-nucleon interaction in medium may be obtained studying the $`K^{}`$-atomic data. The analysis of $`K^{}`$-atomic data using a phenomenological density dependent potential showed that the real part of antikaon optical potential could be as large as $`180\pm 20`$ MeV . It was first demonstrated by Kaplan and Nelson within a chiral $`SU(3)_L\times SU(3)_R`$ model that $`K^{}`$ meson may undergo Bose-Einstein condensation in dense matter formed in heavy ion collisions . The effective mass of antikaons decreases with increasing density because of the strongly attractive $`K^{}`$-baryon interaction in dense matter. Consequently, the in-medium energy of $`K^{}`$ mesons in the zero-momentum state also decreases with density. The $`s`$-wave $`K^{}`$ condensation sets in when the energy of $`K^{}`$ mesons equals to its chemical potential. We consider $`K^{}`$ condensation as a first order phase transition. Here we adopt a relativistic field theoretical model to describe (anti)kaon-baryon interaction . In this case, (anti)kaon-baryon interaction is treated in the same footing as baryon-baryon interaction. The pure $`K^{}`$ condensed phase is composed of members of the baryon octet embedded in the condensate, electrons and muons and maintains charge neutrality and beta-equilibrium conditions. The charge neutrality condition is $`Q^{\overline{K}}=_Bq_Bn_B^{\overline{K}}n_K^{}n_en_\mu =0`$, where $`n_B^{\overline{K}}`$ is the number density of baryon B in pure antikaon condensed phase and $`n_K^{}`$ is the number density of antikaons. Strangeness changing processes such as, $`np+K^{}`$ and $`e^{}K^{}+\nu _e`$, in a neutron star interior are responsible for the onset of $`K^{}`$ condensation. The requirement of chemical equilibrium yields $`\mu _n\mu _p`$ $`=`$ $`\mu _K^{}=\mu _e,`$ (5) where $`\mu _K^{}`$ is the chemical potentials of $`K^{}`$ mesons. The above condition determines the onset of $`K^{}`$ condensation. The deconfinement phase transition from hadronic to quark matter is a possibility in a neutron star interior. In this calculation, we also consider a first order phase transition from hadronic to quark phase. The pure quark matter is composed of $`u`$, $`d`$ and $`s`$ quarks. The EoS of pure quark matter is described in the MIT bag model . As we are considering first order phase transitions, the mixed phase of two pure phases is governed by Gibbs phase rules and global conservation laws . The Gibbs phase rules read, $`P^I=P^{II}`$ and $`\mu _B^I=\mu _B^{II}`$ where $`P^I`$, $`P^{II}`$ and $`\mu _B^I`$, $`\mu _B^{II}`$ are pressure and chemical potentials of baryon B in phase I and II respectively. The conditions for global charge neutrality and baryon number conservation are respectively $`(1\chi )Q^I+\chi Q^{II}=0`$ and $`n_B=(1\chi )n_B^I+\chi n_B^{II}`$ where $`\chi `$ is the volume fraction in phase II. The total energy density in the mixed phase is given by $`ϵ=(1\chi )ϵ^I+\chi ϵ^{II}`$. ## 3 Results & Discussions We adopt GM1 parameter set for this calculation where nucleon-meson coupling constants are determined from the nuclear matter saturation properties. The vector meson coupling constants for (anti)kaons and hyperons are determined from the quark model whereas the scalar meson coupling constants for hyperons and antikaons are obtained from the potential depths of hyperons and antikaons in normal nuclear matter . As the phenomenological fit to the $`K^{}`$ atomic data yielded a very strong real part of antikaon potential $`U_K^{}`$ at normal nuclear matter density, we perform this calculation with an antikaon optical potential of -160 MeV at normal nuclear matter density ($`n_0=0.153fm^3`$). The coupling constants for strange mesons with hyperons and (anti)kaons are taken from Ref.. Also, we consider a bag constant $`B^{1/4}`$ = 200 MeV and strange quark mass $`m_s`$ = 150 MeV for quark matter EoS. We compute equations of state for four different compositions of neutron star matter. The EoS with $`n`$, $`p`$, $`e`$ and $`\mu `$ is denoted by ”np”. With further inclusion of hyperons, it becomes ”npH”. The EoS undergoing a first order phase transition from nuclear to quark matter is represented by ”npQ”. In the last case, the EoS involves first order phase transitions from nuclear to $`K^{}`$ condensed matter and then from $`K^{}`$ condensed matter to quark matter and is referred to as ”npKQ” . It is evident from Eq. (4) that neutron and electron chemical potentials are two independent quantities. One can express energy density and pressure of neutron star matter using $`\mu _n`$ and $`\mu _e`$. As an example, we demonstrate here the construction of EoS involving a first order phase transition from nuclear to $`K^{}`$ condensed matter. In Figure 1, pressure is plotted as a function of neutron and electron chemical potentials for pure hadronic and $`K^{}`$ condensed phase. Here shaded surface denotes the antikaon condensed phase whereas the wired surface represents the hadronic phase. It is noted that two surfaces intersect along a common line. This implies that two phases having same pressure along the intersection line could co-exist. In this figure, the solid line represents the EoS of charge neutral and beta-equilibrated matter. It has three parts. Below P=42 MeV fm<sup>-3</sup>, the pressure in pure hadronic phase is higher than that of the antikaon condensed phase making the hadronic phase physically preferred one. Therefore, the solid line below pressure 42 MeV fm<sup>-3</sup> denotes the charge neutrality line in pure hadronic phase. This phase is mainly composed of neutrons, protons, electrons and muons. Two surfaces first meet at 2.23$`n_0`$ and P=42 MeV fm<sup>-3</sup> and this is the beginning of the mixed phase. After the onset of the phase transition, the mixed phase represented by the solid line follows the intersection part of two surfaces. The Gibbs conditions of phase equilibrium and global charge neutrality are satisfied along the intersection line of two surfaces. We find that the electron chemical potential $`\mu _e`$ decreases in the mixed phase. This is attributed to the fact that with the formation of $`K^{}`$ condensate, electrons are replaced by $`K^{}`$ mesons. The phase transition ends at 3.51$`n_0`$ and pressure 70 MeV fm<sup>-3</sup>. The solid line above pressure 70 MeV fm<sup>-3</sup> represents the charge neutrality line in pure antikaon condensed phase. In Figure 2, the mass-radius relationship of non-rotating neutron stars is displayed for different EoS. For np case, the maximum neutron star mass is 2.39 M and the corresponding radius is 11.94 km. With the inclusion of softening components such as hyperons, Bose-Einstein condensate of $`K^{}`$ mesons or quarks, the maximum neutron star masses are reduced. For npH case, the maximum mass and the corresponding radius are 1.74M and 12.41 km respectively. Maximum masses and the corresponding radii are 1.83 M and 12.88 km for npQ case and 1.75M and 12.10 km for npKQ case, respectively. Among all cases considered here, the maximum neutron star mass for np case is the largest because the EoS in this case is the stiffest. Now we compare our results with the recent findings from EXO 0748-676. Villarreal and Strohmayer have predicted that the best fit values of mass and radius for EXO 0748-676 are M$``$ 1.8 M and R$``$ 11.5 km respectively. It is found that theoretical M-R relationships shown in Figure 2 are consistent with the observed values. It shows that there might be room for exotic matter in neutron star interior. Now we focus on the estimate of radius from the calculation of moment of inertia of pulsar A in the double pulsar binary PSR J0737-3039. In this calculation, we adopt some of EoS used for explaining EXO 0748-676 data. Also, we use the inputs gravitational mass M=1.337M and angular velocity $`\mathrm{\Omega }`$ = 276.8 s<sup>-1</sup> corresponding to a spin period P=$`2\pi /\mathrm{\Omega }`$=22.7 ms to calculate the moment of inertia of the uniformly rotating star. It follows from the calculation that for np, npH and npQ cases the central energy density corresponding to pulsar A is $``$ 5.4 $`\times `$ 10<sup>14</sup> g cm<sup>-3</sup>. We also note that the moment of inertia and radius of pulsar A are I $``$ 4.90 $`\times `$ 10<sup>45</sup> g cm<sup>2</sup> and R $``$ 14 km for all these cases studied here. It exhibits that hyperons or quark matter does not change the moment of inertia and radius appreciably. In Figure 3, we have shown the moment of inertia in dimensionless unit as a function of compactness parameter (M/R) of the neutron star for different EoS. Already, it has been noted that these EoS give rise to maximum neutron star masses $`>`$ 1.7 M. Therefore, $`I/MR^2`$ versus M/R curves shown in Figure 3 can be approximated by $$I(0.237\pm 0.008)MR^2\left[1+4.2\frac{Mkm}{M_{}R}+90\left(\frac{Mkm}{M_{}R}\right)^4\right].$$ (7) The values of moment of inertia that we have obtained from the rotating neutron star model for fixed angular frequency are comparable with the values calculated using Eq. (7). Thus from the measurements of moment of inertia and mass of pulsar A, the radius could be determined by inversion of Eq. (7). It has been shown that the moment of inertia of a neutron star depends quite strongly on the EoS. The predictions for neutron star moment of inertia and radius are quite different in non-relativistic and relativistic models . Therefore, a measurement of moment of inertia with 10 percent uncertainty could impose new constraints on the EoS and could rule out some classes of equations of state that currently exist. ## 4 Summary and conclusions We have computed various properties of static and uniformly rotating compact stars using EoS with first order nuclear to $`K^{}`$ condensed matter phase transition and also $`K^{}`$ condensed matter to quark matter phase transition. Mass and radius of EXO 0748-676 have been measured. These informations put stringent constraints on EoS. Our results are consistent with the mass and radius of EXO 0748-676. This, in turn, shows that bizarre matter might exist in this star. We have also calculated the moment of inertia for pulsar A of double pulsar binary PSR J0737-3039. It is found that the values of moment of inertia and radius obtained with relativistic EoS are quite different from that of the nonrelativistic model of Akmal and Pandharipande . Soon the measurement of moment of inertia for pulsar A would be possible. This would lead to an estimate of the radius of pulsar A. Therefore, it would be possible to rule out some EoS with the help of the measured value of moment of inertia and mass of pulsar A. Lattimer and Prakash showed that there was an empirical correlation between the radius and matter pressure in the density region n<sub>0</sub> to 2n<sub>0</sub> . Therefore, the determination of radius of pulsar A having a mass M=1.337M could give an estimate of matter pressure around saturation density. It is interesting to note that we have informations about mass and radius of a single neutron star. More events like this would be available in future. Therefore, it would be worth investigating the dense matter EoS from the knowledge of masses and radii of neutron stars and using a numerical inversion of neutron star structure equation . Acknowledgments: Two authors (SB and DB) acknowledge the support for this work by the Department of Science and Technology (DST), Government of India and German Academic Exchange Service (DAAD), Germany. ## References
warning/0506/hep-ph0506140.html
ar5iv
text
# Soft Hadron Ratios at LHC ## I Introductory Remarks Relativistic heavy ion collisions experimental program has as objective the formation of the deconfined quark–gluon plasma (QGP) phase in the laboratory. The uncertainty in this experimental program is if, in the available short collision time, $`10^{22}`$$`10^{23}`$ s, the color frozen nuclear phase can melt and turn into the QGP state of matter. There is no valid first principles answer available today, nor it seems, it can be expected in the foreseeable future. From this, and other such uncertainties about the QGP, arises the need to define and study its observables, even though we are quite convinced that this is the state of matter that filled the Universe in its early stage, till hadronization occurred 10–20$`\mu `$s into its evolution. QGP is the equilibrium state in a hot Universe at temperature above that of lightest hadronic particle, the pion, $`kT>m_\pi c^2=140`$ MeV (we hence use units such that $`k=c=\mathrm{}=1`$). Detailed theoretical study of the properties of this new state of matter shows that QGP is rich in entropy and strangeness. These are the observables discussed here explicitly, and implicitly, in the context of soft hadron production. The enhancement of entropy $`S`$ arises in the early stages of the collision process, because the color bonds are broken, and numerous gluons are formed and thermalized. Enhancement of strangeness $`s`$ is in part also due to the breaking of color bonds. Furthermore, it is due to a modification of the kinetic strangeness formation processes. These operate faster in deconfined phase, mainly because the mass threshold for strangeness excitation is considerably lower in QGP than in hadron matter, but also because there are more channels available considering the color quantum numbers. We will not discuss the following important QGP observables in this work, their current status at RHIC is discussed, e.g., in Ref. Huang:2005nd . Aside from strangeness and entropy enhancement, another soft hadron signature is the shape of particle spectra, which carries information about the formation and evolution dynamics of the state of matter that is the source of these particles. Given the considerable increased energy, we expect a greater energy density in the initial stage, and thus a much more violent transverse outflow of matter than has been seen at RHIC. Such a strange transverse collective flow carries many particles to high transverse momenta, and produces a strong azimuthal asymmetry in particle spectra for finite impact parameter reactions. Among other related hadronic signatures, we note a significant charm quark abundance, originating in primary parton reactions. The pattern of charm hadronization should reveal further details about the QGP phase, just as strange hadrons do. Within a few years, a new energy domain will become accessible in study of heavy ion collisions, when the Large Hadron Collider (LHC) at CERN becomes operational in 2007. The top energy available to Pb–Pb reactions is $`\sqrt{s_{\mathrm{NN}}}=5500`$ GeV, a 27.5 fold increase compared to the top RHIC energy. Extrapolating the trends of SPS and RHIC physics, we expect a much greater entropy and strangeness yields at central rapidity. We will always address, in this work, most central head on collision reactions of two $`A200`$ heavy nuclei, at LHC this will be Pb–Pb collision . At these high energies, there will be much less stopping power of baryon number, and thus the central rapidity region will be much more similar to the phase prevailing in the early Universe than this is the case at RHIC. One of the objectives of this work is to assess how small the baryochemical potential $`\mu _\mathrm{B}`$ can become, compared to $`\mu _\mathrm{B}25`$ MeV observed at RHIC. The scale of $`\mu _\mathrm{B}`$ at the hadronization of the early Universe is $`\mu _\mathrm{B}^\mathrm{U}1`$ eV Fromerth:2002wb . In this paper, we address specifically the pattern of soft hadron production based on the assumption of a sudden breakup of the deconfined hadron phase with all soft hadrons produced at essentially same physical conditions, and not subject to requirement of the absolute chemical equilibrium condition. We will, however, provide reference data for the chemical equilibrium case. In the next section II, we outline the statistical hadronization model and present its parameters. In section III, we discuss how simple, but general, hypothesis allow to fix values of these parameters. We establish range of physical interest for strangeness phase space occupancy $`\gamma _s`$. In section IV, we develop our predictions regarding the properties of the hadronizing fireball. We present the range of statistical parameters which we expect and the physical properties of the fireball at its breakup. We consider particle ratios which could help determine the value of $`\gamma _s`$, which is a parameter in this study. In section V, we consider observables sensitive to $`\mu _\mathrm{B}`$, and obtain results which show how precise the hadron yields need to be measured in order to allow measurement of $`\mu _\mathrm{B}`$. ## II Statistical Hadronization and Model Parameters Statistical Hadronization Model (SHM) is, by definition, a model of particle production in which the formation process of each particle fully saturates (maximizes) the quantum mechanical probability amplitude. Particle yields are thus determined by the appropriate integrals of the accessible phase space JJBook . For a system subject to global dynamical evolution, such as collective flow, this applies within each volume element, in its local rest frame of reference. The SHM is consistent with the wealth of SPS and RHIC data available today. Systematic study of particle production for a wide reaction energy range confirms applicability of the SHM, see Letessier:2005qe ; BDMRHIC . Analysis of hadron yields further facilitates a study of the physical properties of the hadronic fireball at the time of hadronization, when these particles are produced, i.e., undergo chemical freeze-out. A study of hadron multiplicities, produced at energy ranging from the top of AGS energy to the top of RHIC energy, leads to an understanding of the physical properties of the fireball at hadronization Letessier:2005qe . Here, we reverse the approach — using the systematics of the energy dependence of the physical properties of the fireball, we establish our expectations about the statistical parameters and relative particle multiplicities expected at LHC. We cannot address, in this work, the total hadron yield, since this depends on the early stage of the reaction, and specifically, on the entropy formation process in initial interactions. The complexity of the SHM model derives from the need to account for many hadronic resonances, and their decays. Since the number of massive resonances grows exponentially, their contributions to particle yields, especially to the yields of pions, are slowly convergent. Moreover, the counting of resonances and their conserved quantum numbers poses a significant book keeping challenge. For this reason, an effort has been made to generate a comprehensive software package available to all interested parties, with a transparent hadron data input, and a comprehensive parameter field. All results presented were obtained using this numerical package SHARE (Statistical Hadronization with REsonances) share . Another package of similar capability, but with restrained parameter set, has since become available Wheaton:2004qb . These programs are including a large number of resonances and track the chemical composition as well as the decay trees with care. As result, the benchmark fits produce a hadronization temperature which is considerably lower than obtained in Ref. BDMRHIC for SPS or RHIC reaction systems. A successful description of rapidity particle yields within the SHM, at a single-chemical freeze-out condition, produces the model parameters in the process of $`\chi ^2`$ minimization. The parameters are: 1) $`dV/dy`$, the volume related a given rapidity to the particle yields; 2) $`T`$, the (chemical) freeze-out temperature; 3) $`\mu _BT\mathrm{ln}(\lambda _u\lambda _d)^{3/2}`$, the baryon and 4) $`\mu _ST\mathrm{ln}[\lambda _q/\lambda _s]`$, hyperon chemical potentials; 5) $`\lambda _{I3}\lambda _u/\lambda _d`$, a fugacity distinguishing the up from the down quark flavor; 6) $`\gamma _s`$ the strangeness phase space occupancy; 7) $`\gamma _q`$ the light quark phase space occupancy. When the assumption of absolute chemical equilibrium is made the values $`\gamma _s=\gamma _q=1`$ are set. The relationship to quark flavor fugacities $`\lambda _i,i=u,d,s`$ is made explicit above. When a set of parameters is known, all particle multiplicities can be evaluated exactly. Similarly, one can obtain the physical properties of the fireball such as thermal energy, entropy, baryon content by appropriate evaluation of the properties of partial fraction contributions of each hadronic state. In that way, in fact, one obtains from a fit to a limited set of measured particle yields a full phase space extrapolation for any particle yield, and a full understanding of the properties of the fireball. Some striking ‘conservation of fireball properties’ rules emerged from such an analysis of particle yield data Letessier:2005qe , and these we will use in order to predict the (relative) particle yields at LHC energy range. ## III LHC Hadronization ### III.1 Choice of conditions and constraints To predict the 7 parameters, at first sight, we need at least 7, and better more than that, valid conditions, constraints and hypothesis: 1) The ‘volume’ normalization $`dV/dy`$ only enters absolute hadron yields. Consequently, it is related to the initial conditions, i.e., mechanisms of entropy production. Restricting our investigation to the study of particle ratios, we do not need to know the value of this parameter, which normalizes the overall yield. There are two natural physics constraints: 2) Strangeness conservation, i.e., the (grand canonical) count of $`s`$ quarks in hadrons equals $`\overline{s}`$ count at each rapidity unit. In our specific case, we request that: $$\frac{\overline{s}s}{\overline{s}+s}=0\pm 0.01.$$ (1) We will show how this condition establishes the relationship between the baryochemical potential $`\mu _\mathrm{B}`$ and the strangeness chemical potential $`\mu _\mathrm{S}`$ in some detail in subsection V.1. 3) The electrical charge to net baryon ratio, in the final state, should be the same as in the initial state, and in the specific case of Pb–Pb interactions, we have: $$\frac{Q}{b}=0.39\pm 0.01.$$ (2) The 2.5% error can be seen as expressing uncertainty about how well neutron and proton densities follow each other, given that even most central reactions occur at finite impact parameter, and 10% of nucleons do not participate in the reaction. 4) and 5) i) The phase space occupancy parameters in some approaches are tacitly fixed: assuming the chemical equilibrium one sets $$\gamma _s=\gamma _q=1.$$ ii) In the chemical non-equilibrium approach, the systematics of data analysis at RHIC and high SPS energies firmly predicts: $$\gamma _qe^{m_{\pi ^\mathrm{o}}/2T}.$$ Furthermore, we will present our results as function of $`\gamma _s`$. We believe that the value of $`\gamma _s`$ is linked to the collision energy, as more strangeness can be produced, when the initial conditions reached become more extreme. We show, in section IV.2, how $`\gamma _s`$ fixes several easily accessible observables and can be measured, and the consistency of the chemical non-equilibrium approach checked. At LHC energy, the expected value of $`\gamma _s`$ is so much greater than unity, and thus, we can be sure that a distinction from $`\gamma _s=1`$ for the equilibrium model can be arrived at in an unambiguous way. There are (at least) two further conditions required which must be sensitive to the hadronization temperature and baryochemical potential. These will be drawn from the following observations: 6) We note the value, $$\frac{E}{TS}0.78,\mathrm{or}0.845,$$ for chemical non-equilibrium Letessier:2005qe , or respectively, equilibrium model analysis. We note that the energy per particle of non-relativistic and semi-relativistic classical particle gas is $`E/Nm+3/2T+\mathrm{}`$, while the entropy per particle in this condition is $`S/N=5/2+m/T+\mathrm{}`$ (see section 10 of JJBook ). Hence: $$\frac{E}{TS}\frac{m/T+3/2}{m/T+5/2}.$$ (3) It is thus possible to interpret this constraint in terms of a quark matter made of particles with thermal mass $`maT`$. Solving Eq. 3, we find for $`E/TS=0.78`$, $`a=2`$ for chemical non-equilibrium, and for $`E/TS=0.845`$, $`a=4`$ for equilibrium. This is near to the result expected in finite temperature QCD Petreczky:2001yp . That result points to a simple structure of the quark matter fireball. On the quark-side, the value $`E/TS`$ is not very model specific, though it is sensitive to the average particle mass as shown above. On the other hand, there is considerable sensitivity to this thermodynamic constraint in the hadronic gas. The hadron system after hadronization comprises a mix of particles of different, and for the baryon component, large mass. To fine tune this value a specific ratio of baryons to mesons needs to be established: in this way the hadron system can maintain both energy and entropy, aside of baryon number and strangeness, during hadronization. For this reason, in the hadron phase there is considerable sensitivity of this ratio to both $`T`$ and the value of the phase space occupancy parameters, here $`\gamma _s`$. 7) We need to make an assumption which fixes the baryochemical potential $`\mu _\mathrm{B}`$. This certainly is one of most difficult guesses as there is no reliable way to predict baryon stopping at LHC, and certainly this value, $$\mu _\mathrm{B}T,$$ will be quite difficult to measure. For this reason, we will discuss, in subsection V at length a method to measure $`\mu _\mathrm{B}`$. The impact of $`\mu _\mathrm{B}1`$–3 MeV on mixed particle ratios, such as $`\mathrm{K}/\pi `$ or $`\eta /\pi ^0`$ is physically irrelevant. We will show haw particle–antiparticle abundances can help us further. In fact, it is uncertain that a measurement of such small $`\mu _\mathrm{B}`$ can be accomplished, and thus in effect, we could have simply assumed $`\mu _\mathrm{B}=0`$ which at a few %-level would be consistent with all relative hadron yield predictions here presented. On the other hand, the understanding of matter–antimatter asymmetry present at LHC energy scale is, in itself, of interest and thus, we pursue this question further. We argue as follows: our analysis of RHIC data suggested that baryons are more easily retained in the central rapidity region than energy, with 2.5 times larger fraction of colliding baryons than fraction of energy deposited. Considering the SPS data point, and the RHIC results, the per baryon thermalized reaction energy retained in the central rapidity region, in units of the maximum available collision energy, drops from 40% available at RHIC to 15% at LHC. Given the LHC energy $`\sqrt{s_{\mathrm{NN}}}=5500`$ GeV, a thermal energy content per net baryon at $`y=0`$ is assumed to be, $$\frac{dE}{db}=0.15\times 5500/2=412\pm 20\mathrm{GeV},$$ at top LHC energy, where the error is chosen to be similar in relative magnitude as the error in other observables considered. This error plays a role in finding the solution in terms of statistical variables of the constraints and conditions considered. This specific assumption fixes implicitly the (small) value of the baryochemical potential, and by virtue of the strangeness conservation, also of the strange chemical potential. When we deviate from this assumption in exploring a wider parameter range, we will mention this explicitly. ### III.2 $`\eta `$ and maximum value of $`\gamma _s`$ We will study the hadronization condition as function of $`\gamma _s`$ which may take large values. We note that $`\gamma _s`$ cannot rise above a limit, to be determined from similar consideration as is the maximum value of $$\gamma _q^{\mathrm{CR}}=\mathrm{exp}(m_\pi /2T).$$ (4) At this value, the Bose distributions of pions diverges. As $`\gamma _s`$ increases, same will happen in the strange hadron sector and indeed this will occur first to the lightest particle with considerable hidden strangeness content, i.e., $`\eta (548)`$. Naively, one could expect that $`\gamma _s^{\mathrm{CR}}=\mathrm{exp}(m_\eta /2T)`$. However, $`\eta `$, unlike the spin 1 $`\varphi (1020)`$ is not a (nearly) pure $`\overline{s}s`$ state. The quark structure of $`\eta (548)`$ and $`\eta ^{}(958)`$ can be written as: $`\eta `$ $`=`$ $`{\displaystyle \frac{u\overline{u}+d\overline{d}}{\sqrt{2}}}\mathrm{cos}\varphi _ps\overline{s}\mathrm{sin}\varphi _p,`$ (5) $`\eta ^{}`$ $`=`$ $`{\displaystyle \frac{u\overline{u}+d\overline{d}}{\sqrt{2}}}\mathrm{sin}\varphi _p+s\overline{s}\mathrm{sin}\varphi _p.`$ Study of numerous experimental results, and in particular of $`Z^0`$ hadrons, shows that $`\mathrm{sin}^2\varphi _p=0.45\pm 6`$ indicating that $`\eta (548)`$ has 45% strangess content Uvarov:2001wv . This arises from the SU(3)-flavor octet state content of 67% reduced by SU(3)-symmetry breaking mixing with the 33% strangeness content of the singlet $`\eta ^{}(958)`$. In order to count the yield of the $`\eta `$, we introduce its fugacity $`\mathrm{{\rm Y}}_\eta `$. The fractional contribution to the partition function is: $$\mathrm{ln}Z_\eta =𝑑V\frac{d^3p}{(2\pi )^3}\mathrm{ln}\left(1\mathrm{{\rm Y}}_\eta e^{\frac{E_\eta }{T}}\right),$$ (6) with $`E_\eta =\sqrt{m_\eta ^2+p^2}`$. We focus our attention on the dominant, directly produced $`\eta `$. The incremental per unit volume $`\eta `$ yield is: $$\frac{dN_\eta }{dV}=\mathrm{{\rm Y}}_\eta \frac{[d\mathrm{ln}Z_\eta /dV]}{\mathrm{{\rm Y}}_\eta }=\frac{d^3p}{(2\pi )^3}\frac{1}{\mathrm{{\rm Y}}_\eta ^1e^{\frac{E_\eta }{T}}1}.$$ (7) More specifically, the rapidity density is: $$n_\eta \frac{dN_\eta }{dy}=\frac{dV}{dy}\times \frac{dN_\eta }{dV}.$$ (8) The ‘normalization’ $`dV/dy`$ which comprises the transverse dimension at hadronization, and the longitudinal incremental volume, is arising from kinetic expansion processes driven by the initial state formation mechanisms. These have to be obtained by methods beyond the scope of this work. However, we note that in the longitudinally scaling limit for ideal fluid hydrodynamic evolution of the initial state, at all times the entropy rapidity density remains constant, $`dS/dy=\mathrm{Const}.`$. Since the particle multiplicity is defined by the value of $`dS/dy`$, the total hadron multiplicity is not affected by our ensuing study of hadronization for different chemical freeze-out temperatures. The total charge particle rapidity density is a consequence of initial processes which are beyond the physics reach of this work. We now relate the $`\eta `$-fugacity $`\mathrm{{\rm Y}}_\eta `$ to the light and strange quark fugacities $`\gamma _q`$ and $`\gamma _s`$. In the SHM, the probability of the production of $`\eta `$ is weighted with the yield of strange and light quark pairs in proportion of their contribution to the quark content in the particle formed. Thus, $$\mathrm{{\rm Y}}_\eta =\gamma _q^2\mathrm{cos}^2\varphi _p+\gamma _s^2\mathrm{sin}^2\varphi _p<e^{m_\eta /T}.$$ (9) The upper limit is set at the phase space divergence point. Specifically, considering $`\eta (548)`$ with its 45% strangeness content, we find from Eq. (9) a maximal value $`\gamma _s<10.4`$ for the hadronization conditions of interest in this work, i.e., $`\gamma _qe^{m_\pi /2T}`$ and $`T140`$ MeV. Thus we set as the range of interest $`0<\gamma _s<10`$. This upper limit is the most stringent constraint for $`\gamma _s`$: it is more stringent than the one which arises from the consideration of the $`\varphi `$-meson, $$\mathrm{{\rm Y}}_\varphi =\gamma _s^2<e^{m_\varphi /T},$$ (10) due to the greater $`\varphi `$-mass. Similarly, the constraint on $`\gamma _s`$ based on the kaon condensation (in presence of negligible chemical potentials): $$\mathrm{{\rm Y}}_\mathrm{K}=\gamma _s\gamma _q<e^{m_\mathrm{K}/T}.$$ (11) is less severe. To confirm the functional dependence on $`\gamma _i`$, Eq. (9), we show that it is consistent with the requirement that the fugacity $`\gamma _i`$, allows the count of valance quark content in hadrons. We look at $`\gamma _s`$, which allows the count of all strange and antistrange quarks by the relation: $$s+\overline{s}=\gamma _s\frac{\mathrm{ln}Z}{\gamma _s}.$$ (12) The contribution of $`\eta `$ to the strangeness count thus is, $`(s+\overline{s})_\eta `$ $`=`$ $`{\displaystyle \frac{\gamma _s}{\mathrm{{\rm Y}}_\eta }}{\displaystyle \frac{\mathrm{{\rm Y}}_\eta }{\gamma _s}}\mathrm{{\rm Y}}_\eta {\displaystyle \frac{\mathrm{ln}Z_\eta }{\mathrm{{\rm Y}}_\eta }},`$ (13) $`=`$ $`2{\displaystyle \frac{\gamma _s^2\mathrm{sin}^2\varphi _p}{\gamma _q^2\mathrm{cos}^2\varphi _p+\gamma _s^2\mathrm{sin}^2\varphi _p}}N_\eta ,`$ which is the required result. A similar procedure can be followed to show that: $$(q+\overline{q})_\eta =2\frac{\gamma _q^2\mathrm{cos}^2\varphi _p}{\gamma _q^2\mathrm{cos}^2\varphi _p+\gamma _s^2\mathrm{sin}^2\varphi _p}N_\eta .$$ (14) We conclude that Eq. (9) defines the $`\eta `$-fugacity in terms of light and strange quark fugacities. ### III.3 Range of $`\gamma _s`$ of interest When $`\gamma _s`$ increases, the $`\eta `$-fugacity increases rapidly, but, even at $`\gamma _s=5`$, it is well below the Bose singularity. For $`T=160`$ MeV, the singularity would be just above $`\gamma _s=8`$. However, the fall-off of the expected hadronization temperature (see below) moves this towards twice as large value. One may wonder how large a value of $`\gamma _s`$ is physically consistent, and in particular, if an associate decrease in the value of $`T`$ makes sense. In qualitative terms, this type of parameter evolution and correlation is fully consistent with the picture of rapid transverse expansion of the QGP fireball. This expansion can lead to supercooling which pushes the hadronization temperature lower. At the same time, the preserved yield of strangeness requires that $`\gamma _s`$ increases. We now look at the possible range of expected values of $`\gamma _s`$ in more detail. It can be expected that, for LHC extreme conditions, the strangeness phase space has been chemically saturated at a temperature larger than strange quark mass: $`T_1>m_s`$. Moreover, in the QGP phase, there is a residual strange quark mass $`m_s>T`$, where $`T`$ is the final hadronization temperature. While the precise value of $`m_s`$ depends on the momentum scale at which it is measured, we will assume in this semi-quantitative discussion that $`m_s(1\text{ GeV})180`$ MeV. Conservation of strangeness rapidity yield in the expansion from $`T_1`$ to $`T`$ implies, by comparison of the relativistic phase space size, $$(dV_1T_1^3)W(m_s/T_1)=\gamma _s^\mathrm{Q}\times (dVT^3)W(m_s/T),$$ (15) where $`\gamma _s^\mathrm{Q}`$ is the quark-side hadronization phase space occupancy, for definition of $`W(x)`$, see Eq. (18). For $`0<x<1`$, $`W(x)`$ is slowly changing and close to its asymptotic value $`W(x=0)=2`$. Entropy conservation further requires that $`dV_1T_1^3=dVT^3=\text{Const.}`$ For $`T_11.5T`$, and noting that $`T_1>m_s>T`$, we obtain, as solution of Eq. (15), $`\gamma _s^\mathrm{Q}1.5`$. In the hadronization process, this value increases by a significant factor which we can obtain comparing the phase space of strange hadrons in QGP with that hadron phases (we omit as is customary the upper index H, i.e., $`\gamma _s^\mathrm{H}\gamma _s`$): $`\gamma _s^\mathrm{Q}\times W(m_s/T)`$ $`=`$ $`\gamma _s\times {\displaystyle \underset{i}{}}\stackrel{~}{W}(m_i/T)`$ $`+`$ $`\gamma _s^2\times {\displaystyle \underset{j}{}}\stackrel{~}{W}(m_j/T)`$ $`+`$ $`\gamma _s^3\times {\displaystyle \underset{k}{}}\stackrel{~}{W}(m_k/T).`$ The sums run over single, double and triply strange ($`s`$ and $`\overline{s}`$) hadrons, respectively. $`\stackrel{~}{W}`$ differs from $`W`$ in that it comprises appropriate hadron fugacities for each particle, the appropriate expressions are seen in more detail in Eq. (17). The strangeness QGP to HG aspect ratio, i.e., the ratio of the phase phase space size seen in Eq. (III.3) up to the coefficients $`\gamma _s^\mathrm{Q}`$, $`\gamma _s`$ shows that it grows with decreasing temperature, see Fig. 19.3 in Ref. JJBook . This is further strongly amplifying by a factor as large as four the final observed $`\gamma _s`$. A QGP phase value $`\gamma _s^\mathrm{Q}1.5`$ may become $`\gamma _s5`$–10 on hadron side. As we shall see, the decrease of $`T`$ with increasing $`\gamma _s`$ compensates, in part, the increase with $`\gamma _s`$ in the yields of strange hadrons. We will thus consider the range $`0.5<\gamma _s<10`$ as that is where most of variation in the considered observables is seen, and the actual physical conditions are expected to occur. The lower limit underscores the comparison with the equilibrium model $`\gamma _s=\gamma _q=1`$ behavior. The upper limit $`\gamma _s<10`$ is within the range of allowed values of $`\gamma _s`$ we have obtained in previouse section III.2. ## IV Predictions ### IV.1 SHM parameter values at LHC top energy With the assumptions outlined in subsection III.1, we solve for the best set of SHM model parameters. A precise solution is always found in the considered range of $`\gamma _s`$, also for the increased value of $`E/TS`$ for the case of chemical equilibrium. This of course does not guarantee that, e.g., our baryochemical potential is correctly chosen. The results we show are not entirely smooth as we allow a small error in the constraints and conditions, and thus the solution for the parameters is not a precise algebraic result, but a most likely value of a quasi-fit, which has a $`\chi ^20`$. The resulting statistical parameters $`T`$, $`\gamma _q^{\mathrm{CR}}`$, $`\mu _\mathrm{B}`$, and $`\mu _\mathrm{S}`$ are shown, in Fig. 1, as function of $`0.5\gamma _s10`$. Note that through our study of chemical non-equilibrium $`\gamma _q\gamma _q^{\mathrm{CR}}`$. We see that an increasing value of $`\gamma _s`$ is accompanied by considerable reduction of the hadronization temperature, which drops from the value $`T=140`$ MeV, near $`\gamma _s=2.4`$ and $`\gamma _q=1.6`$, down to $`T=110`$ MeV. At the favorite value $`\gamma _s=5`$, the expected LHC hadronization temperature is $`T=125`$ MeV for the chemical non-equilibrium. The chemical equilibrium result at $`T=156`$ implies a chemical potential $`\mu _\mathrm{B}=2.6`$ MeV, and this value is also found for the chemical non-equilibrium model for $`\gamma _s=5`$. We explain why $`\mu _\mathrm{B}`$ remains unchanged at the end of section V. For $`\mu _\mathrm{S}`$, we see a slight reduction by 10% from equilibrium model value at 0.52 MeV as is appropriate considering the results presented in Fig. 5. Our estimate of the chemical parameters at LHC are considerably different from those proposed by others Andronic:2003zv , where $`\mu _\mathrm{B}=1`$ MeV and $`\mu _\mathrm{S}=0.3`$ MeV is proposed. We note that our baryochemical potential is considerably greater. We will discuss experimental consequences at greater length in section V. It is of considerable interest to study the physical properties at hadronization: the pressure $`P`$, the energy density $`ϵ=E/V`$, entropy density $`\sigma =S/V`$, and net baryon density $`\nu =B/V`$. In general, the non-equilibrium hadronization occurs from a state of higher density, as is seen in Fig. 2, comparing the lines with the SHM equilibrium cross. This is, in particular, true for the entropy density. With increasing $`\gamma _s`$, all density decreases, the drop in temperature is a more important influence than the large increase in relative strangeness yield. Despite a significant increase in $`\mu _\mathrm{B}`$ with increasing $`\gamma _s`$, the net baryon density decreases modestly. It is very small, bordering the value $`\nu =0.001`$ fm<sup>-3</sup>. ### IV.2 Particle yield ratios and determination of $`\gamma _s`$ There is approximate charge symmetry with positives $`h^+`$ and negatives $`h^{}`$ having a very similar yield. The difference will be discussed further below. The total charged hadron yield will be denoted as, $$h=h^++h^{}p+\overline{p}+\pi ^++\pi ^{}+\mathrm{K}^++\mathrm{K}^{},$$ and is evaluated after weak decay of hyperons and $`\mathrm{K}_{\mathrm{S},\mathrm{L}}`$. Similarly the total yield of neutrals: $$h^0\pi ^0+n+\overline{n}.$$ The top panel, in Fig. 3, shows the $`2h^0/h`$ ratio, which varies by $`\pm 10`$% in the range of $`\gamma _s`$ considered, with the charge symmetric value $`2h^0/h=1`$ arising at $`\gamma _s=3.3`$. In the lower three panels, we focus our interest on ratios of some interest for determination of $`\gamma _s`$. We present the ratio $`\eta /\pi ^0`$ which rises by nearly 50% compared to the equilibrium model expectation, see crosses at $`\gamma _s=1`$, computed for $`\gamma _q=1`$ and $`T=156`$ MeV. This ratio is observed by reconstruction of the invariant di-photon mass. This observable shows a relatively small variation with $`\gamma _s`$, which can be understood as result of competition between $`\gamma _s`$ and $`\gamma _q`$, see Eq. (9), which is accompanied by the decrease of $`T`$ with increasing $`\gamma _s`$. We record that the expectation for the non-equilibrium hadronization at LHC is: $$0.07<\frac{\eta }{\pi ^0}<0.12.$$ This interesting observable may also not have the sensitivity required to distinguish the models, or help determine $`\gamma _s`$. In the panel below, we show the $`(\mathrm{K}^++\mathrm{K}^{})/h`$ ratio. This ratio is rising, for large $`\gamma _s`$, to a value near 0.23 almost double the ‘standard’ chemical equilibrium value at 0.12. However, for $`\gamma _s=5`$, we see a more modest increase by 50%. We record for LHC: $$0.12<\frac{\mathrm{K}^++\mathrm{K}^{}}{h^++h^{}}<0.23.$$ More spectacular is the expected increases in the $`2\varphi /h`$ ratio. The Chemical equilibrium value of 0.015 is seen to rise 4-fold, and at $`\gamma _s=5`$, we still see a very noticeable increase by a factor 2.5. This is a very important observable of the condition of hadronization. We record our LHC expectation: $$0.015<\frac{2\varphi }{h^++h^{}}<0.06.$$ The behavior of the baryon yields is shown in Fig. 4 where the total, nearly matter–antimatter symmetric yields of protons, and the three hyperon families are shown, normalized by the total charged hadron yield $`h`$. The lines, from top to bottom are for $`(p+\overline{p})/h`$, $`(\mathrm{\Lambda }+\overline{\mathrm{\Lambda }})/h`$, $`(\mathrm{\Xi }^{}+\overline{\mathrm{\Xi }}^+)/h`$ and $`(\mathrm{\Omega }^{}+\overline{\mathrm{\Omega }}^+)/h`$. We note that for large $`\gamma _s`$, a considerable change in the expected baryon populations ensues, with proton yield decreasing (due to decrease in $`T`$ while the more strange the hyperon is, the more it is enhanced compared to chemical equilibrium expectations. Interestingly, these relative yields saturate with increasing $`\gamma _s`$, as the effect of temperature decrease competes with the increase due to rising $`\gamma _s`$. Note that we have shown, in Fig. 4, the total hadron yields after all weak decays have occurred. We summarize our LHC expectations: $`0.07>`$ $`{\displaystyle \frac{p+\overline{p}}{h^++h^{}}}`$ $`>0.04,`$ $`0.02<`$ $`{\displaystyle \frac{\mathrm{\Lambda }+\overline{\mathrm{\Lambda }}}{h^++h^{}}}`$ $`<0.04,`$ $`0.004<`$ $`{\displaystyle \frac{\mathrm{\Xi }^{}+\overline{\mathrm{\Xi }}^+}{h^++h^{}}}`$ $`<0.015,`$ $`0.0006<`$ $`{\displaystyle \frac{\mathrm{\Omega }^{}+\overline{\mathrm{\Omega }}^+}{h^++h^{}}}`$ $`<0.004.`$ ## V Measurement of the baryo-chemical potential ### V.1 Strangeness conservation An interesting challenge, at LHC, will be the measurement of chemical potentials. We recall that $`\mu _\mathrm{S}`$ is directly related to $`\mu _\mathrm{B}`$ should hadron emission at each rapidity occurs such that there is local strangeness balance, i.e., ‘conservation’. This relates the two chemical potentials as we shall next discuss. Otherwise, if emission of hadrons were to reflect on the QGP conditions, we would expect $`\mu _\mathrm{S}=0`$ and this would generate a buildup of residual strangeness in a distillation process destil . We tacitly assumed, and continue in this way now, that at LHC, in each rapidity region, local conservation of strangeness prevails. We present a set of results which will allow, given appropriate experimental sensitivity, to determine the value of $`\mu _\mathrm{S}`$. Strangeness conservation establishes a relation between the chemical potentials ChemCons . While this relationship is simplified for the case $`\mu _i/T1`$, the presence of $`\gamma _s1`$ introduces new elements and we reinspect the relationship. Using, at first, the quark fugacity notation for convenience, the open strangeness sector partition function is, in the here appropriate Boltzmann approximation: $`\mathrm{ln}Z_s`$ $``$ $`\gamma _s\gamma _qF_\mathrm{K}\left({\displaystyle \frac{\lambda _s}{\lambda _q}}+{\displaystyle \frac{\lambda _q}{\lambda _s}}\right)+\gamma _s\gamma _q^2F_\mathrm{Y}\left(\lambda _s\lambda _q^2+{\displaystyle \frac{1}{\lambda _s\lambda _q^2}}\right)`$ (17) $`+\gamma _s^2\gamma _qF_\mathrm{\Xi }\left(\lambda _s^2\lambda _q+{\displaystyle \frac{1}{\lambda _s^2\lambda _q}}\right)+\gamma _s^3F_\mathrm{\Omega }\left(\lambda _s^3+{\displaystyle \frac{1}{\lambda _s^3}}\right).`$ The phase space factors, $$F_i(T)=\frac{VT^3}{2\pi ^2}\underset{ki}{}g_kW(m_k/T),W(x)=x^2K_2(x),$$ (18) comprise all contributing hadron states ‘$`k`$’ with quantum number ‘$`i`$’. $`W(x)=x^2K_2(x)\stackrel{x0}{}2`$ is the relativistic phase space integral in classical limit, which for large $`x`$ behaves as $`W(x)x^{3/2}\mathrm{exp}(x/T)`$. A series of these terms represents quantum statistics, see Eq. (V.1). The strangeness conservation condition, $$s\overline{s}=\frac{\lambda _s\mathrm{ln}Z}{\lambda _s}=0,$$ yields for small values of chemical potentials, in units of $`T`$: $`\mu _\mathrm{B}\mu _\mathrm{S}+2(\mu _\mathrm{B}2\mu _\mathrm{S}){\displaystyle \frac{\gamma _s}{\gamma _q}}{\displaystyle \frac{F_\mathrm{\Xi }}{F_\mathrm{Y}}}+3(\mu _\mathrm{B}3\mu _\mathrm{S}){\displaystyle \frac{\gamma _s^2}{\gamma _q^2}}{\displaystyle \frac{F_\mathrm{\Omega }}{F_\mathrm{Y}}}`$ $`=`$ $`\mu _\mathrm{S}\left({\displaystyle \frac{F_\mathrm{K}}{\gamma _qF_\mathrm{Y}}}+{\displaystyle \frac{\gamma _s\stackrel{~}{F}_\mathrm{K}(2m_\mathrm{K})}{4F_\mathrm{Y}}}+{\displaystyle \frac{\gamma _q\gamma _s^2\stackrel{~}{F}_\mathrm{K}(3m_\mathrm{K})}{9F_\mathrm{Y}}}+\mathrm{}\right).`$ The right hand side presents the three first terms of the kaon Bose integral expansion which keeps terms of the same order as those belonging to $`\mathrm{\Xi }`$ and $`\mathrm{\Omega }`$ in the balance. These are not entirely negligible for large $`\gamma _s`$. We solve Eq. (V.1) and show, in Fig. 5, $`\mu _\mathrm{B}/\mu _\mathrm{S}`$ as function of $`T`$: $$\frac{\mu _\mathrm{B}}{\mu _\mathrm{S}}=f(T;\gamma _q,\gamma _s).$$ (20) The dashed curve is the equilibrium case for $`\gamma _s=\gamma _q=1`$, where we marked the RHIC hadronization condition $`\mu _\mathrm{B}/\mu _\mathrm{S}=5`$ with a cross. The closest solid line to this result is for $`\gamma _q=e^{m_{\pi ^0}/T}`$ and $`\gamma _s0`$. The following solid lines are, in sequence from upper right, for $`\gamma _s=0.5,\mathrm{\hspace{0.17em}1},\mathrm{\hspace{0.17em}3},\mathrm{\hspace{0.17em}5},\mathrm{\hspace{0.17em}7},\mathrm{\hspace{0.17em}10}`$. The cross near the $`\gamma _s=3`$ line, at $`\mu _\mathrm{B}/\mu _\mathrm{S}=4.6`$, corresponds to the RHIC chemical non-equilibrium hadronization condition. These results allow, aside of gauging the LHC values of chemical potentials, also an easy comparison with and cross check of other work addressing LHC and RHIC environments with conserved strangeness. We recall that many particle ratios are directly determined by chemical potentials, e.g., $`\mathrm{K}^+/K^{}=\mathrm{exp}(2\mu _\mathrm{S}/T)`$ and $`\overline{\mathrm{\Xi }}^+/\mathrm{\Xi }^{}=\mathrm{exp}((2\mu _\mathrm{B}4\mu _\mathrm{S})/T)`$. This obviously leads to consistency conditions in the particle–antiparticle asymmetry sensitive particle ratios. ### V.2 Hadron–antihadron asymmetry In order to measure chemical potentials, we need to be able to measure the particle–antiparticle asymmetry. This is not an easy task at LHC as we shall see. Up to RHIC energy, it was customary to study ratio of antiparticle yields to particle yields, such as $`\overline{\mathrm{\Lambda }}/\mathrm{\Lambda }`$. At LHC, the strategy has to slightly change. We consider of normalized particle–antiparticle difference yields: $$\mathrm{\Delta }N_i=\frac{\overline{n}_in_i}{\overline{n}_i+n_i},$$ (21) where $`n_i`$ is the rapidity density $`dN/dy`$ of charged particles $`i=\mathrm{K}^+,p,\mathrm{\Lambda },\mathrm{\Xi },\mathrm{\Omega }`$ and antiparticles. We omitted intentionally the pion from this list, as it is the dominant component of the unidentified particle hadron asymmetry, we will address next. Below, we will omit the factor $`d/dy`$ as the expressions we state are valid more generally. The most accessible observable, $$\mathrm{\Delta }h\frac{h^+h^{}}{h^++h^{}},$$ (22) composed of unidentified charged hadrons is very hard to measure precisely, for i) there are distortion possible by partial acceptance of weak decays, and ii) this variable assumes a comparatively small value at LHC. Moreover, this is also an observable hardest to interpret in a simple and transparent theoretical model. We find that its magnitude for LHC will be: $$\mathrm{\Delta }h=(0.5\text{}0.6)10^3,$$ for all values of $`\gamma _s1`$, as is shown in Fig. 6. The difference in $`\mathrm{\Delta }h`$ between equilibrium and non-equilibrium result is due to the increased entropy and thus hadron multiplicity content of the chemical non-equilibrium case. The variable $`\mathrm{\Delta }h`$ has been recognized already in the study of SPS reactions as a sensitive probe of entropy production ChemCons ; entro . Should a way to measure $`\mathrm{\Delta }h`$ of this magnitude with reasonable precision be found at LHC, this would provide a model independent measure of the value of the baryo-chemical potential $`\mu _\mathrm{B}`$. To see this, we solve (fit) the SHM with a fixed assumed value of $`\mathrm{\Delta }h`$. In the chemical equilibrium version of SHM, we find, at $`\gamma _s=\gamma _q=1`$, in the $`T`$$`\mu _\mathrm{B}`$ plane the constraint lines determined by assumed values of $`\mathrm{\Delta }h`$, shown in the top panel of Fig. 7. From left to right, we have considered three benchmark values, $`\mathrm{\Delta }h=0.001`$ (top LHC energy), $`\mathrm{\Delta }h=0.01`$ (top RHIC energy), and $`\mathrm{\Delta }h=0.1`$ (top SPS energy, central rapidity region). At SPS, top energy in S–Pb interactions, we have studied this variable (called $`D_Q0.009`$) in the context of the exploration of the specific entropy per baryon ChemCons ; entro , and the results we present here are in agreement with this early study of SPS hadron multiplicities. For non-equilibrium case, with $`\gamma _q=\gamma _q^{\mathrm{CR}}`$, the corresponding result is shown in the second top panel. The nearly vertical lines are solution of the the following conditions imposed: strangeness conservation, charge-to-baryon ratio and $`\mathrm{\Delta }h`$, as function of $`\mu _\mathrm{B}`$. For each value of $`T`$, there is a specific value of $`\gamma _s`$ indicated in the next lower panel, and again, we see nearly a vertical line. Both these results imply that while the value of $`T`$ and $`\gamma _s`$ are not much constrained by the measurement of just one single observable $`\mathrm{\Delta }h`$, the value of $`\mu _\mathrm{B}`$ is already highly constrained. Why this is the case is understood inspecting the bottom panel in Fig. 7. We show the resulting value of thermal energy per baryon $`E/b`$. These turn out to be highly localized regions. Thus, $`\mathrm{\Delta }h`$ is for a wide range of other statistical parameters closely related to the value of energy per baryon, or equivalently, entropy per baryon. The connecting dashed line in the bottom panel guides the eye. A check of E/TS also confirms that these solutions produce the expected result which is otherwise introduced as a constraint. Thus, we learn that just the single ‘measured’ value $`\mathrm{\Delta }h`$ is enough to constrain a rather narrow range of $`\mu _\mathrm{B}`$. This result is indicated by the vertical dotted line, which we place bracing the domains of $`\mu _\mathrm{B}`$ that are allowed. Inspecting the equilibrium model intercept, we realize that this singles out a domain of $`T`$ which is result of data fits with this constraint. This explains why $`\mu _\mathrm{B}`$ is usually determined in a model independent way within the SHM, with little if any difference present between the different model variants, provided that the experimental data used in the fit comprises explicitly, or implicitly, $`\mathrm{\Delta }h`$. For example, the study of the impact parameter dependence at RHIC using different SHM model variants produced $`\mu _\mathrm{B}`$ and $`\mu _\mathrm{S}`$ which cannot be distinguished (see bottom panel of Fig. 1 in Ref. bdepend ). ### V.3 Identified particle–antiparticle asymmetries For the identified particles, the normalized particle–antiparticle differences can be closely and analytically related to the value of chemical potential. For example, the kaon asymmetry is directly related to strangeness chemical potential $`\mu _\mathrm{S}`$: $$\mathrm{\Delta }\mathrm{K}\frac{\mathrm{K}^+\mathrm{K}^{}}{\mathrm{K}^++\mathrm{K}^{}}\mathrm{tanh}\frac{\mu _\mathrm{S}}{T}\frac{\mu _\mathrm{S}}{T}.$$ (23) We have, for simplicity, not considered the $`\varphi `$-meson decay contributions which increase the normalizing yield but do not alter the difference. We show the actual, with all decays, $`\mathrm{\Delta }\mathrm{K}`$ as thick solid line in the top panel of Fig. 8 (bottom line in this top panel). The very top short dashed line in the panel (red on-line) is the ratio $`\mu _\mathrm{S}/T`$. The thick long-dashed line excludes from the ratio the contamination by the decay $`\varphi \mathrm{K}^++\mathrm{K}^{}`$. The fully weak decay contamination corrected results are the thin (solid and resp. dashed) lines at the top of the parallel lines, and are shown for both the full result (solid lines) and $`\varphi `$-decay corrected result (long dashed). The parallel line regions are where the acceptance of weak decays is partial and/or the correction is incomplete. After the removal of the $`\varphi `$-decay dilution of the kaon yields, Eq. (23) should read: $$\mathrm{\Delta }\mathrm{K}0.9\frac{\mu _\mathrm{S}}{T}.$$ (24) The slight reduction from the analytical formula Eq. (23) is due to the strong decay contributions of hyperon resonances decaying emitting a kaon. Because of the smallness of $`\mathrm{\Delta }K`$, the baryon asymmetry in the hyperon resonances leaves this small but visible imprint of this result. The chemical equilibrium result (cross before and circle after weak decays) is also indicated in Fig. 8. These are, in general, larger than the analytical results (short dashed lines) except in the case of $`\mathrm{\Delta }K`$. For baryons there are four particle–antiparticle differentials, which are shown below $`\mathrm{\Delta }K`$, in Fig. 8. We expect for protons: $$\mathrm{\Delta }p\frac{p\overline{p}}{p+\overline{p}}=\mathrm{tanh}\frac{\mu _\mathrm{B}}{T}\frac{\mu _\mathrm{B}}{T}.$$ (25) The thin solid line in the second panel from the top which corresponds to removed weak decays compares well to the analytical results, short-dashed line. The thick solid line at bottom of parallel lines includes in $`\mathrm{\Delta }p`$ the contamination from weak decays of $`\mathrm{\Lambda }`$ and $`\overline{\mathrm{\Lambda }}`$, and the region in between spans all possible WD contamination. Once the weak decay contributions $`\mathrm{\Omega },\mathrm{\Xi }\mathrm{\Lambda }`$ and $`\overline{\mathrm{\Omega }},\overline{\mathrm{\Xi }}\overline{\mathrm{\Lambda }}`$ are removed, we have further: $$\mathrm{\Delta }\mathrm{\Lambda }\frac{\mathrm{\Lambda }\overline{\mathrm{\Lambda }}}{\mathrm{\Lambda }+\overline{\mathrm{\Lambda }}}=\mathrm{tanh}\frac{\mu _\mathrm{B}\mu _\mathrm{S}}{T}\frac{\mu _\mathrm{B}\mu _\mathrm{S}}{T}.$$ (26) The thin line, in the $`\mathrm{\Delta }\mathrm{\Lambda }`$ panel, is nearly indistinguishable from this result. The thick solid line includes all weak decays. There is no significant contamination of $`\mathrm{\Xi }^{}`$ and $`\overline{\mathrm{\Xi }}^+`$ and thus we have: $$\mathrm{\Delta }\mathrm{\Xi }\frac{\mathrm{\Xi }^{}\overline{\mathrm{\Xi }}^+}{\mathrm{\Xi }^{}+\overline{\mathrm{\Xi }}^+}=\mathrm{tanh}\frac{\mu _\mathrm{B}2\mu _\mathrm{S}}{T}\frac{\mu _\mathrm{B}2\mu _\mathrm{S}}{T},$$ (27) and similarly for the $`\mathrm{\Omega }^{}`$: $$\mathrm{\Delta }\mathrm{\Omega }\frac{\mathrm{\Omega }^{}\overline{\mathrm{\Omega }}^+}{\mathrm{\Omega }^{}+\overline{\mathrm{\Omega }}^+}=\mathrm{tanh}\frac{\mu _\mathrm{B}3\mu _\mathrm{S}}{T}\frac{\mu _\mathrm{B}3\mu _\mathrm{S}}{T},$$ (28) both shown in bottom panel of Fig. 8. For baryons, the expected asymmetry is at %-level and the weak decay of hyperons allow unique identification of these particles. It is quite possible that the measurement of these variables will succeed. As the above expressions show, there is a consistency condition since two chemical potentials and the temperature considering that the isospin bath of numerous pions causes $`\lambda _{I3}1`$ to a great precision, such that $`T\mathrm{ln}\lambda _{I3}\mu _\mathrm{S}`$ determine five observables $`\mathrm{\Delta }N`$. This consistency is further tightened due to strangeness conservation relation of $`\mu _\mathrm{B}/\mu _\mathrm{S}=4`$–5, see section V.1. The strangeness conservation constraint and through it the SHM model can be tested by transforming Eqs. (23,25,26,27,28) so we can make use of Eq. (V.1) and the values Eq. (20): $`{\displaystyle \frac{\mathrm{\Delta }p}{\mathrm{\Delta }\mathrm{K}}}`$ $`=`$ $`{\displaystyle \frac{\mu _\mathrm{B}}{\mu _\mathrm{S}}}4.5,`$ (29) $`{\displaystyle \frac{\mathrm{\Delta }\mathrm{\Lambda }}{\mathrm{\Delta }\mathrm{K}}}`$ $`=`$ $`\left({\displaystyle \frac{\mu _\mathrm{B}}{\mu _\mathrm{S}}}1\right)3.5,`$ (30) $`{\displaystyle \frac{\mathrm{\Delta }\mathrm{\Xi }}{\mathrm{\Delta }\mathrm{K}}}`$ $`=`$ $`\left({\displaystyle \frac{\mu _\mathrm{B}}{\mu _\mathrm{S}}}2\right)2.5,`$ (31) $`{\displaystyle \frac{\mathrm{\Delta }\mathrm{\Omega }}{\mathrm{\Delta }\mathrm{K}}}`$ $`=`$ $`\left({\displaystyle \frac{\mu _\mathrm{B}}{\mu _\mathrm{S}}}3\right)1.5.`$ (32) These relation test the SHM and strangeness conservation, they do not differentiate model variants such as chemical equilibrium and chemical non-equilibrium. ## VI Final remarks The total multiplicity yield, as well as the yield of charmed particles, is originating predominantly in the early stage, primary parton reactions. For this reason, we did not address absolute yields of hadrons, and, similarly, cannot study the total charm yield in the context considered here. However, one may wonder, if the appearance of small but distinguished charmed meson and baryon yield does not offer an interesting and independent probe of the properties of the hadronization state, or even it could influence the results presented. A earlier study of charmed hadron production, in chemical equilibrium at $`T=170`$ MeV, has yielded interesting insights into the relative production strengths of charmed mesons and baryons emerging at this particular hadronization temperature Andronic:2003zv . We will not enter into further discussion of this subject, but note that: a) for $`\gamma _s>2,`$ these results imply that $`D_s(c\overline{s})^+`$ and its antiparticle should be the dominant charmed hadron fraction; b) the differences in yields between particles and antiparticles, seen in table 3 of Ref. Andronic:2003zv , indicate that charm particle contribution to the asymmetries we studied are totally negligible. It appears that measurement of relative yield of charmed mesons and baryons will reveal the charm hadronization condition, and we hope to return to this subject soon. The small relative number of charmed quarks and their even smaller particle–anti-particle asymmetry assures that these particles do not impact any of the results we obtained. On the other hand, these particles offer another opportunity to explore hadronization conditions. The formation of charmed hadrons is expected to occur prior to general hadronization, considering the greater binding of charmed particles Thews:2004sh . To summarize, we have presented a detailed study of the soft hadron production pattern at LHC. We discussed, in turn, relative yields such as $`\varphi /h`$ and $`K/h`$ which allow insights into the hadronization conditions and help address questions related of chemical equilibrium and non-equilibrium, as well as temperature of hadron freeze-out. Perhaps the most interesting result to pursue experimentally is the large value of $`\gamma _s`$ expected in hadronization of over-saturated QGP phase. The ratios, shown in Figs. 3 and 4, are not very sensitive to the choices we made that determine chemical potentials $`\mu _\mathrm{B}`$ and $`\mu _\mathrm{S}`$, they probe primarily the interplay between the $`\gamma _s`$ and $`T`$. We note further that most these ratios, at the favored value $`\gamma _s5`$, differ considerably from the chemical equilibrium model expectations. The relatively large value of $`\gamma _s`$ we expect at LHC, twice as large as our analysis finds at RHIC, derives from a larger absolute density of strangeness at hadronization of the deconfined phase, combined with lower prevailing temperature expected in deeper expansion supercooling. Indeed, while at RHIC $`\gamma _s^{\mathrm{QGP}}|_{\mathrm{hadronization}}1`$, at LHC we expect $`\gamma _s^{\mathrm{QGP}}|_{\mathrm{hadronization}}1.5`$$`2`$. The difference in the phase space size of QGP with HG than leads to $`\gamma _s5`$. We have further presented an in depth discussion of particle–antiparticle asymmetries which address the challenge of chemical and strange quark chemical potential measurement. Since the particle–antiparticle yield difference is small compared to each individual yield, a special effort will need to be made to acquire these difference ratios $`\mathrm{\Delta }N_i`$ 0.1–4% at the level of 10% or better. Work supported by a grant from the U.S. Department of Energy DE-FG02-04ER41318. LPTHE, Univ. Paris 6 et 7 is: Unité mixte de Recherche du CNRS, UMR7589.
warning/0506/cond-mat0506428.html
ar5iv
text
# 1 Introduction ## 1 Introduction In sharp contrast with the zero magnetic field case, the analysis of properties of electrons in periodic or random potentials subjected to external magnetic fields is a very challenging problem. The difficulty is rooted in the singular nature of the magnetic interaction: due to a linear increase of the magnetic vector potential, the naive perturbation theory breaks down even at arbitrarily small fields. To our best knowledge, only the periodic case has been considered in connection with the Faraday effect for bulk systems. The first full scale quantum computation was done by Laura M. Roth (for a review of earlier attempts we direct the reader to this paper). The physical experiment starts by sending a monochromatic light wave, parallel to the $`0z`$ direction and linearly polarized in the plane $`x0z`$. When the light enters the material, the polarization plane can change; in fact, there exists a linear relation between the angle $`\theta `$ of rotation of the plane of polarization per unit length and the transverse component of the conductivity tensor $`\sigma _{xy}`$ (see formula (1) in ). The material is chosen in such a way that when the magnetic field is zero, this transverse component vanishes. When the magnetic field $`B`$ is turned on, the transverse component is no longer zero. For weak fields one expands the conductivity tensor to first order in $`B`$ and obtains a formula for the Verdet constant. Therefore the central object is $`\sigma _{xy}(B)`$, which depends among other things on temperature, density of the material, and frequency of light. Using a modified Bloch representation, Roth was able to obtain a formula for $`\frac{d\sigma _{xy}}{dB}(0)`$, and studied how this first order term behaves as a function of frequency, both for metals and semiconductors. However, the theory in is not free of difficulties. First, it seems almost hopeless to estimate errors or to push the computation to higher orders in $`B`$. Second, even the first order formula contains terms which are singular at the crossings of the Bloch bands. Accordingly, at the practical level this theory only met a moderate success and alternative formalisms have been used, as for example the celebrated Kohn-Luttinger effective many band Hamiltonian (see and references therein), or tight-binding models . Since all these methods have limited applicability, a more flexible approach was still needed. In the zero magnetic field case, a very successful formalism (see e.g. and references therein) is to use the Green function method. This is based on the fact that the traces involved in computing various physical quantities can be written as integrals involving Green functions. The main aim of our paper is to develop a Green function approach to the Faraday effect, i.e. for the conductivity tensor when a magnetic field is present. Let us point out that the use of Green functions (albeit different from the ones used below) goes back at least to Sondheimer and Wilson in their theory of diamagnetism of Bloch electrons. Aside from the fact that the Green function (i.e. the integral kernel of the resolvent or the semi-group) is easier to compute and control, the main point is that by factorizing out the so called ”non integrable phase factor” (or ”magnetic holonomy”) from the Green function, one can cope with the singularities introduced by the increase at infinity of the magnetic vector potentials. In addition, (as it has already been observed by Schwinger in a QED context) after factorizing out the magnetic holonomy one remains with a gauge invariant quantity which makes the problem of gauge fixing irrelevant. The observation (going back at least to Peierls ) that one can use these magnetic phases in order to control the singularity of the magnetic perturbation has been used many times in various contexts (see e.g. ). We highlight here the results of Nedoluha where a Green function approach for the magneto-optical phenomena at zero temperature and with the Fermi level in a gap has been investigated. But the power of this method has only recently been fully exploited in , and developed as a general gauge invariant magnetic perturbation theory in . Applied to the case at hand, this theory gives an expansion of the conductivity tensor in terms of the zero magnetic field Green functions. Moreover, it is free of any divergences. A key ingredient in controlling divergences is the exponential decrease of the Green functions with the distance between the arguments, for energies outside the spectrum . We stress the fact that since no basis is involved, periodicity is not needed and the theory can also be applied to random systems. Finite systems and/or special geometries (layers) are also allowed. The content of the paper is as follows: In Section 2 we give a derivation of the conductivity tensor from first principles in the linear response theory. We include it to point out that it coincides with various formulas used before. Although in physics establishing the Kubo formula is considered somehow a triviality, from a mathematical point of view it remains a serious chalenge (see ). Section 3 contains the precise formulation of the thermodynamic limit, stated in Theorem 3.1. We do not give its proof here, but we try to explain why it is true. Section 4 shows that at the thermodynamic limit, at zero temperature, zero frequency, and for the Fermi energy in a spectral gap, we re-obtain a formula of Streda for the transverse component of the conductivity tensor, known from the Integral Quantum Hall Effect (IQHE). A precise statement and its proof are contained in Theorem 4.1. Moreover, under the proviso that exponentially localized Wannier function exist (see Theorem 4.2), this transverse component vanishes (see also , for related results). We stress that this result holds for the whole $`\sigma _{xy}(B)`$ as long as the magnetic field is not too large, not just for $`\frac{d\sigma _{xy}}{dB}(0)`$. The vanishing of its first order correction was in fact claimed in formula (50) in . Section 5 contains the exact quantum computation of $`\sigma _{xy}(B)`$ for free electrons; in spite of the fact that such a result might be known (and it is known at zero frequency), we were not able to find it in the literature. Interestingly enough, the quantum computation gives the same result as the well known classical computation (when the relaxation time is infinite). Section 6 contains the core of the paper, which includes the derivation of $`\frac{d\sigma _{xy}}{dB}(0)`$ for general Bloch electrons. As in the zero magnetic field case, its formula only contains zero magnetic field Green functions and current operators. Section 7 deals with periodic systems, and the result of the previous section is written down in terms of zero magnetic field Bloch functions and bands. At the end we have some conclusions. The main goal of this paper is to present the strategy, state the results concerning the Verdet constant, and to outline future theoretical and practical problems. Detailed proofs of the thermodynamic limit and of other technical estimates will be given elsewhere. ## 2 Preliminaries. The conductivity tensor in the linear response regime We begin by fixing the notation used in the description of independent electrons subjected to a constant magnetic field. The units are chosen so that $`\mathrm{}=1`$. Since we consider spin $`1/2`$ particles, the one particle Hilbert space for a non-confined particle is $$_{\mathrm{}}=L^2(^3)L^2(^3)$$ with the standard scalar product. Accordingly, all operators below and their integral kernels are $`2\times 2`$ matrices in the spin variable. We choose the constant magnetic field of strength $`B`$ to be oriented along the $`z`$-axis. Then the one particle Hamiltonian with the spin-orbit coupling included is (see e.g. ) $$H_{\mathrm{}}(B)=\frac{1}{2m}𝐏(B)^2+V+g\mu _bB\sigma _3,$$ (2.1) with $$𝐏(B)=ib𝐚+\frac{1}{2mc^2}𝐬(V)=𝐏(0)b𝐚$$ (2.2) where $$b=\frac{e}{c}B$$ and $`𝐚(𝐱)`$ is an arbitrary smooth magnetic vector potential which generates a magnetic field of intensity $`B=1`$ i.e. $`𝐚(𝐱)=(0,0,1)`$. The most frequently used magnetic vector potential is the symmetric gauge: $$𝐚_0(𝐱)=\frac{1}{2}𝐧_3𝐱.$$ (2.3) where $`𝐧_3`$ is the unit vector along $`z`$ axis. In the periodic case we denote by $``$ the underlying Bravais lattice, by $`\mathrm{\Omega }`$ its elementary cell and by $`\mathrm{\Omega }^{}`$ the corresponding Brillouin zone. $`|\mathrm{\Omega }|`$ and $`|\mathrm{\Omega }^{}|`$ stand for the volumes of the elementary cell and Brillouin zone respectively. In the absence of the magnetic field one has the well known Bloch representation in terms of Bloch functions: $$\mathrm{\Psi }_j(𝐱,𝐤)=\frac{1}{\sqrt{|\mathrm{\Omega }^{}|}}e^{i𝐤𝐱}u_j(𝐱,𝐤),𝐱^3$$ (2.4) where $`u_j(𝐱,𝐤)`$ are the normalized to one eigenfunctions of the operator $$h(𝐤)u_j(𝐱,𝐤)=\lambda _j(𝐤)u_j(𝐱,𝐤)$$ (2.5) $`h(𝐤)`$ $`={\displaystyle \frac{1}{2m}}\left(i_p+{\displaystyle \frac{1}{2mc^2}}𝐬(V)+𝐤\right)^2+V,`$ $`={\displaystyle \frac{1}{2m}}(𝐩+𝐤)^2+V,𝐤\mathrm{\Omega }^{},`$ (2.6) $`𝐩`$ $`=i_p+{\displaystyle \frac{1}{2mc^2}}𝐬(V),`$ defined in $`L^2(\mathrm{\Omega })L^2(\mathrm{\Omega })`$ with periodic boundary conditions. We label $`\lambda _j(𝐤)`$ in increasing order. We have to remember that, as functions of $`𝐤`$, $`\lambda _j(𝐤)`$ and $`u_j(𝐱,𝐤)`$ are not differentiable at the crossing points. Since the $`\mathrm{\Psi }_j(𝐱,𝐤)`$’s form a basis of generalized eigenfunctions, the Green function (i.e. the integral kernel of the resolvent) writes as: $$G_{\mathrm{}}^{(0)}(𝐱,𝐲;z)=_\mathrm{\Omega }^{}\underset{j1}{}\frac{|\mathrm{\Psi }_j(𝐱,𝐤)\mathrm{\Psi }_j(𝐲,𝐤)|}{\lambda _j(𝐤)z}d𝐤,$$ (2.7) and it is seen as a matrix in the spin variables. The above formula has to be understood in the formal sense since the series in the right hand side is typically not absolutely convergent, and care is to be taken when interchanging the sum with the integral. Notice however that $`G_{\mathrm{}}^{(0)}(𝐱,𝐱^{};z)`$ is a well behaved matrix valued function. We consider a system of noninteracting electrons in the grand-canonical ensemble. More precisely, we consider a box $`\mathrm{\Lambda }_1^3`$, which contains the origin, and a family of scaled boxes $$\mathrm{\Lambda }_L=\{𝐱^3:𝐱/L\mathrm{\Lambda }_1\}.$$ (2.8) The thermodynamic limit will mean $`L\mathrm{}`$, that is when $`\mathrm{\Lambda }_L`$ tends to fill out the whole space. The one particle Hilbert space is $`_L:=L^2(\mathrm{\Lambda }_L)L^2(\mathrm{\Lambda }_L)`$. The one particle Hamiltonian is denoted by $`H_L(B)`$ and is given by (2.1) with Dirichlet boundary conditions (i.e. the wave-functions in the domain of $`H_L(B)`$ vanish at the surface $`\mathrm{\Lambda }_L`$). More precisely, we first define it on $`C_0^{\mathrm{}}(\mathrm{\Lambda }_L)C_0^{\mathrm{}}(\mathrm{\Lambda }_L)`$, and then $`H_L(B)`$ will be the Friedrichs extension of this minimal operator. This is indeed possible, because our operator can be written as (up to some irrelevant constants) $`\mathrm{\Delta }_DI_2+W,`$ where $`\mathrm{\Delta }_D`$ is the Dirichlet Laplacian and $`W`$ is a first order differential operator, relatively bounded to $`\mathrm{\Delta }_DI_2`$ (remember that $`L<\mathrm{}`$) with relative bound zero. The form domain of $`H_L(B)`$ is the Sobolev space $`H_0^1(\mathrm{\Lambda }_L)H_0^1(\mathrm{\Lambda }_L)`$, while the operator domain is $$\mathrm{Dom}(H_L(B))=D_LD_L,D_L:=H^2(\mathrm{\Lambda }_L)H_0^1(\mathrm{\Lambda }_L).$$ (2.9) Moreover, $`H_L(B)`$ is essentially self-adjoint on $`C_{(0)}^{\mathrm{}}(\overline{\mathrm{\Lambda }_L})`$, i.e. functions with support in $`\overline{\mathrm{\Lambda }_L}`$ and indefinitely differentiable in $`\mathrm{\Lambda }_L`$ up to the boundary. We assume that the temperature $`T=1/(k\beta )`$ and the chemical potential $`\mu `$ are fixed by a reservoir of energy and particles. We work in a second quantized setting with an antisymmetric Fock space denoted by $`_L`$. Denote the operators in the Fock space with a hat and borrow some notation from the book of Bratelli and Robinson : if $`A`$ is an operator defined in $`_L`$, we denote by $`\widehat{A}=d\mathrm{\Gamma }(A)`$ its second quantization in the Fock space. At $`t=\mathrm{}`$ the system is supposed to be in the grand-canonical equilibrium state of temperature $`T`$ and chemical potential $`\mu `$, i.e. the density matrix is $$\widehat{\rho }_e=\frac{1}{\mathrm{Tr}(e^{\beta \widehat{K}_\mu })}e^{\beta \widehat{K}_\mu },$$ (2.10) where $$\widehat{K}_\mu =d\mathrm{\Gamma }(H_L(B)\mu \mathrm{Id})$$ (2.11) is the “grand-canonical Hamiltonian”. The interaction with a classical electromagnetic field is described by a time dependent electric potential $$V(𝐱,t):=(e^{i\omega t}+e^{i\overline{\omega }t})e𝐄𝐱,t0,𝐱\mathrm{\Lambda }_L.$$ (2.12) so the total time dependent one-particle Hamiltonian is $$H(t)=H_L(B)+V(t).$$ (2.13) Notice that $`e`$ near $`𝐄`$ is the positive elementary charge. Here we take $`\mathrm{Im}\omega <0`$ which plays the role of an adiabatic parameter, and insures that there is no interaction in the remote past. Finally, the one-particle current operator is as usual $$𝐉=ei[H_L(B),𝐗]=\frac{e}{m}𝐏(B),$$ (2.14) where $`𝐗`$ is the multiplication by $`𝐱`$. Note that J is a well defined operator on the domain of $`H_L(B)`$, because multiplication by any component of $`𝐗`$ leaves this domain invariant (see (2.9)). Moreover, since $`L<\mathrm{}`$, $`𝐗`$ is a bounded operator. In fact, $`𝐗`$ is the true physical self-adjoint observable, while $`𝐏(B)`$ (or $`𝐉`$) appear when one differentiates the map $`te^{itH_L(B)}`$X$`e^{itH_L(B)}`$ in the strong sense on the domain of $`H_L(B)`$. We assume that the state of our system is now described by a time-dependent density matrix, $`\widehat{\rho }(t)`$, obtained by evolving $`\widehat{\rho }_e`$ from $`\mathrm{}`$ up to the given time, i.e. $$i_t\widehat{\rho }(t)=[\widehat{H}(t),\widehat{\rho }(t)],\widehat{\rho }(\mathrm{})=\widehat{\rho }_e.$$ (2.15) Going to the interaction picture and using the Dyson expansion up to the first order, one gets $$\widehat{\rho }(t=0)=\widehat{\rho }_ei_{\mathrm{}}^0[d\mathrm{\Gamma }(\stackrel{~}{V}(s),\widehat{\rho }_e]ds+𝒪(𝐄^2),$$ (2.16) where $$\stackrel{~}{V}(s):=e^{isH_L(B)}V(s)e^{isH_L(B)}.$$ (2.17) The current density flowing through our system at $`t=0`$ is given by (see (2.16)): $`𝐣`$ $`={\displaystyle \frac{1}{|\mathrm{\Lambda }_L|}}\mathrm{Tr}__L\left(\widehat{\rho }(0)\widehat{𝐉}\right)={\displaystyle \frac{1}{|\mathrm{\Lambda }_L|}}\mathrm{Tr}__L\left(\widehat{\rho }_e\widehat{𝐉}\right)`$ $`{\displaystyle \frac{i}{|\mathrm{\Lambda }_L|}}\mathrm{Tr}__L\left({\displaystyle _{\mathrm{}}^0}[d\mathrm{\Gamma }(\stackrel{~}{V}(s)),\widehat{\rho }_e]\widehat{𝐉}𝑑s\right)+𝒪(𝐄^2).`$ (2.18) In evaluating the r.h.s. of (2) we use the well known fact that traces over the Fock space can be computed in the one-particle space (see Proposition 5.2.23 in ): $$\mathrm{Tr}__L\left\{\widehat{\rho }_ed\mathrm{\Gamma }(A)\right\}=\mathrm{Tr}__L\left\{f_{FD}(H_L(B))A\right\},$$ (2.19) where $`f_{FD}`$ is the Fermi-Dirac one-particle distribution function: $$f_{FD}(x):=\frac{1}{e^{\beta (x\mu )}+1},x,\beta >0,\mu .$$ (2.20) Plugging (2.19) into (2), the identity $`[d\mathrm{\Gamma }(A),d\mathrm{\Gamma }(B)]=d\mathrm{\Gamma }([A,B])`$, the invariance of trace under cyclic permutations and ignoring the quadratic correction in $`𝐄`$ one arrives at $`𝐣`$ $`={\displaystyle \frac{1}{|\mathrm{\Lambda }_L|}}\mathrm{Tr}__L\left\{f_{FD}(H_L(B))𝐉\right\}`$ $`{\displaystyle \frac{i}{|\mathrm{\Lambda }_L|}}{\displaystyle \frac{e}{m}}\mathrm{Tr}__L\left({\displaystyle _{\mathrm{}}^0}[\stackrel{~}{V}(s),𝐏(B)]f_{FD}(H_L(B))𝑑s\right).`$ (2.21) The first term in (2) is always zero because of the identity (trace cyclicity again) $$\mathrm{Tr}__L\left\{[H_L(B),𝐗]f_{FD}(H_L(B))\right\}=\mathrm{Tr}__L\left\{[f_{FD}(H_L(B)),H_L(B)]𝐗\right\}=0.$$ (2.22) which is nothing but the fact that the current vanishes on an equilibrium state. Note that these operations under the trace sign are quite delicate, since unbounded operators are involved. Let us for once give a complete proof to (2.22). We have the identity between bounded operators (consider the first component $`X_1`$): $$[H_L(B),X_1]f_{FD}(H_L(B))=H_L(B)X_1f_{FD}(H_L(B))X_1H_L(B)f_{FD}(H_L(B)).$$ (2.23) Remember that $`X_1`$ is a bounded operator in the box, and preserves the domain of $`H_L(B)`$. This means that the operator $`O_L=(H_L(B)+i)X_1(H_L(B)+i)^1`$ is bounded. Hence we can write $$H_L(B)X_1f_{FD}(H_L(B))=[1i(H_L(B)+i)^1]O_L[H_L(B)+i]f_{FD}(H_L(B)).$$ Now the operator $`[H_L(B)+i]f_{FD}(H_L(B))`$ still is trace class due to the exponential decay of $`f_{FD}`$, while $`[1i(H_L(B)+i)^1]`$ and $`O_L`$ are bounded. Thus $`H_L(B)X_1f_{FD}(H_L(B))`$ is trace class and we can compute its trace using the complete eigenbasis of $`H_L(B)`$, which gives the same result as for the other operator $`X_1H_L(B)f_{FD}(H_L(B))`$. Thus (2.22) is proved. Using (2.12) and (2.17) one can write $`j_\alpha ={\displaystyle \underset{\beta =1}{\overset{3}{}}}\{\sigma _{\alpha \beta }(\omega )+\sigma _{\alpha \beta }(\overline{\omega })\}E_\beta ,\alpha \{1,2,3\},\mathrm{}(\omega )<0,`$ (2.24) where the conductivity tensor is given by $`\sigma _{\alpha \beta }(B,\omega )=`$ (2.25) $`{\displaystyle \frac{i}{|\mathrm{\Lambda }_L|}}{\displaystyle \frac{e^2}{m}}\mathrm{Tr}__L{\displaystyle _{\mathrm{}}^0}[e^{isH_L(B)}x_\beta e^{isH_L(B)},P_\alpha (B)]f_{FD}(H_L(B))e^{is\omega }𝑑s.`$ Performing an integration by parts, using the formulas $`i[H_L(B),x_\beta ]=P_\beta (B)/m`$ and $`i[P_\alpha (B),x_\beta ]=\delta _{\alpha \beta }`$ one arrives at $`\sigma _{\alpha \beta }(B,\omega )={\displaystyle \frac{1}{|\mathrm{\Lambda }_L|}}{\displaystyle \frac{e^2}{im\omega }}\{\delta _{\alpha \beta }\mathrm{Tr}(f_{FD}(H_L(B)))`$ $`+`$ $`{\displaystyle \frac{i}{m}}\mathrm{Tr}{\displaystyle _{\mathrm{}}^0}e^{is(\omega +H_L(B))}P_\beta (B)e^{isH_L(B)}[P_\alpha (B),f_{FD}(H_L(B))]ds\},`$ and this coincides (at least at the formal level) with formula (5) in . Notice that from now on, we write just Tr when we perform the trace, since we only work in the one-particle space. Since we are interested in the Faraday effect, and we assume that the magnetic field $`𝐁`$ is parallel with the $`z`$ axis, we will only consider the transverse conductivity $`\sigma _{12}(B,\omega )`$. Hence the first term vanishes. We now perform the integral over $`s`$ with the help of Stone’s formula followed by a deformation of the contour (paying attention not to hit the singularities of $`f_{FD}(z)`$ or to make the integral over $`s`$ divergent $$f_{FD}(H_L(B))e^{is(H_L(B)+\eta )}=\frac{i}{2\pi }_{\mathrm{\Gamma }_\omega }f_{FD}(z)e^{is(z+\eta )}(H_L(B)z)^1𝑑z.$$ (2.27) where $`\eta `$ is either $`0`$ or $`\omega `$, the contour is counter-clockwise oriented and given by $$\mathrm{\Gamma }_\omega =\{x\pm id:ax<\mathrm{}\}\{a+iy:dyd\}$$ (2.28) with $$d=\mathrm{min}\{\frac{\pi }{2\beta },\frac{|\mathrm{Im}\omega |}{2}\},$$ (2.29) and $`a+1`$ lies below the spectrum of $`H_L(B)`$. As a final result one gets $`\sigma _{12}(B,\omega )={\displaystyle \frac{e^2}{2\pi m^2\omega |\mathrm{\Lambda }_L|}}`$ (2.30) $`\mathrm{Tr}{\displaystyle _{\mathrm{\Gamma }_\omega }}f_{FD}(z)\{P_1(B)(H_L(B)z)^1P_2(B)(H_L(B)z\omega )^1`$ $`+zz\omega \}dz=:{\displaystyle \frac{e^2}{m^2\omega }}a_L(B,\omega )`$ where “$`zz\omega `$” means a similar term where we exchange $`z`$ with $`z\omega `$. Now one can see that by inserting the eigenbasis of $`H_L(B)`$ one obtains the well known formula derived from semi-classical radiation theory (see e.g. formula (4) in ). ## 3 Gauge invariance and existence of the thermodynamic limit Up to now the system was confined in a box $`\mathrm{\Lambda }_L`$. As is well known (see e.g. ) a direct evaluation of (2.30) (or previous formulas equivalent to it including formula (4) in Roth’s paper) is out of reach: the eigenvalues and eigenstates of $`H(B)`$ are rather complicated (even in the thermodynamic limit $`\mathrm{\Lambda }_L^3`$) and at the same time the Bloch representation is plagued by singular matrix elements of the magnetic vector potential. Roth used a modified magnetic Bloch representation in and derived a formula for the linear term in $`B`$ of (2.30) in terms of the zero magnetic field Bloch representation. Still, her procedure is not free of difficulties since it involves $`_𝐤u_j(𝐱,𝐤)`$ which might not exist at crossing points. In addition, it seems almost hopeless to control the errors or to push computations to the second order in $`B`$ which would describe the Cotton-Mouton effect for example. In what follows, we shall outline another route of evaluating (2.30) which is mathematically correct, systematic, and completely free of the above difficulties. There are two basic ideas involved. The first one (going back at least to Sondheimer and Wilson in their theory of diamagnetism) consists in writing the trace in (2.30) as integrals over $`\mathrm{\Lambda }_L`$ of corresponding integral kernels. This is nothing but the well known Green function approach (see e.g. ) which has been very successful in computing optical and magneto-optical properties of solids (see e.g. , , ) in the absence of an external magnetic field. The point is that the integral kernels are on one hand easier to control and compute, and on the other hand they do not require periodicity. Moreover, this approach proved to be essential in deriving rigorous results concerning the diamagnetism of free electrons and actually we expect the methods of the present paper to simplify the theory of diamagnetism of Bloch electrons as well. However, when applying Green function approach in the presence of an external magnetic field one hits again the divergences caused by the linear increase of the magnetic vector potential: naively, at the first sight $`a_L(B,\omega )`$ is not bounded in the thermodynamic limit $`L\mathrm{}`$ but instead grows like the second power of $`L`$. It was already observed in that these divergent terms vanish identically due to some identities coming from gauge invariance. This is indeed the case and the main point of this paper is to show, following the developments in , , , that factorizing the so called “ non-integrable phase factor” from the Green function (the integral kernel of $`(H_L(B)\zeta )^1`$) allows, at the same time, to eliminate the divergences coming from the increase of the magnetic vector potential and to obtain a controlled expansion in powers of $`B`$. In addition, this leads to expressions of $`a_L(B,\omega )`$ which are manifestly gauge invariant. For an arbitrary pair of points $`𝐱`$, $`𝐲\mathrm{\Lambda }_L`$ consider the “magnetic phase” associated with the magnetic vector potential $`𝐚(𝐮)`$ defined as the path integral on the line linking $`𝐲`$ and $`𝐱`$: $$\varphi _𝐚(𝐱,𝐲)=_𝐲^𝐱𝐚(𝐮)𝑑𝐮.$$ (3.1) The magnetic phase satisfies the following crucial identity: for every fixed $`𝐜`$ $`e^{ib\varphi _𝐚(𝐱,𝐜)}𝐏(B)e^{ib\varphi _𝐚(𝐱,𝐜)}=𝐏(0)b𝐀(𝐱𝐜).`$ (3.2) where $`𝐀(𝐱)=\frac{1}{2}𝐧_3𝐱`$, i.e. irrespective of the choice of $`𝐚(𝐱)`$, $$𝐀(𝐱𝐜)=\frac{1}{2}𝐧_3(𝐱𝐜)$$ (3.3) is the symmetric (transverse, Poincaré) gauge with respect to $`𝐜`$. Write now the Green function (as a $`2\times 2`$ matrix in the spin space) $$G_L(𝐱,𝐲;\zeta )=(H_L\zeta )^1(𝐱,𝐲)$$ (3.4) in the factorized form $$G_L(𝐱,𝐲;\zeta )=e^{ib\varphi _𝐚(𝐱,𝐲)}K_L(𝐱,𝐲;\zeta ).$$ (3.5) It is easy to check that while $`G_L(𝐱,𝐲;\zeta )`$ is gauge dependent, $`K_L(𝐱,𝐲;\zeta )`$ is gauge independent i.e. the whole gauge dependence of $`G_L(𝐱,𝐲;\zeta )`$ is contained in the phase factor $`e^{ib\varphi _𝐚(𝐱,𝐲)}`$. Plugging the factorization (3.5) into the integrand of the r.h.s. of (2.30), using (3.2) and (3.3), one obtains that its integral kernel writes as $`𝒜_{s,s^{}}^L(𝐱,𝐱^{})=e^{ib\varphi _𝐚(𝐱,𝐱^{})}`$ (3.6) $`{\displaystyle _{\mathrm{\Gamma }_\omega }}dzf_{FD}(z){\displaystyle \underset{\sigma =1}{\overset{2}{}}}{\displaystyle _{\mathrm{\Lambda }_L}}d𝐲e^{ib\mathrm{\Phi }(𝐱,𝐲,𝐱^{})}\{[(P_{1,𝐱}(0)bA_1(𝐱𝐲))K_L(𝐱,𝐲;z)]_{s,\sigma }`$ $`[(P_{2,𝐲}(0)bA_2(𝐱𝐲))K_L(𝐲,𝐱^{};z+\omega )]_{\sigma ,s^{}}+zz\omega \},`$ where $$\mathrm{\Phi }(𝐱,𝐲,𝐱^{})=\varphi _𝐚(𝐱,𝐲)+\varphi _𝐚(𝐲,𝐱^{})+\varphi _𝐚(𝐱^{},𝐱)$$ is the flux of the magnetic field $`(0,0,1)`$ through the triangle $`\mathrm{\Delta }(𝐱,𝐲,𝐱^{})`$. Now the fact that there are no long range divergences in the formula for $`𝒜_{s,s^{}}(𝐱,𝐱^{})`$ follows from the exponential decay of Green functions (see also ): for $`\zeta `$ outside the spectrum of $`H`$ there exists $`m(\zeta )>0`$ such that as $`|𝐱𝐲|\mathrm{}`$ $$|K_L(𝐱,𝐲;\zeta )|=|G_L(𝐱,𝐲;\zeta )|e^{m(\zeta )|𝐱𝐲|}.$$ (3.7) It can be proved (the technical details which are far from being simple will be given elsewhere) that $`𝒜_{s,s^{}}^L(𝐱,𝐱^{})`$ is jointly continuous and moreover outside a thin region near the surface of $`\mathrm{\Lambda }_L`$ one can replace it by the integral kernel $`𝒜_{s,s^{}}^{\mathrm{}}(𝐱,𝐱^{})`$ of the corresponding operator on the whole $`^3`$. Accordingly, up to surface corrections: $$a_L(B,\omega )\frac{1}{2\pi |\mathrm{\Lambda }_L|}\underset{s=1}{\overset{2}{}}_{\mathrm{\Lambda }_L}𝒜_{s,s}^{\mathrm{}}(𝐱,𝐱)𝑑𝐱.$$ (3.8) Notice that due to the fact that $`\mathrm{\Phi }(𝐱,𝐲,𝐱)=\varphi _𝐚(𝐱,𝐱)=0`$ the phase factors appearing in (3.6) reduce to unity in (3.8). In the periodic case, from the fact that in the symmetric gauge the Hamiltonian $`H_{\mathrm{}}(B)`$ commutes with the magnetic translations (actually one can define magnetic translations for an arbitrary gauge, just first make the gauge transformation relating $`𝐚(𝐱)`$ to $`𝐀(𝐱)`$) generated by $``$, it follows that for $`\stackrel{}{\gamma }`$ we have: $$K_{\mathrm{}}(𝐱+\stackrel{}{\gamma },𝐲+\stackrel{}{\gamma };\zeta )=K_{\mathrm{}}(𝐱,𝐲;\zeta ),$$ which implies that $$𝒜_{s,s}^{\mathrm{}}(𝐱+\stackrel{}{\gamma },𝐱+\stackrel{}{\gamma })=𝒜_{s,s}^{\mathrm{}}(𝐱,𝐱)$$ is periodic with respect to $``$, hence up to surface corrections: $$a_L(B,\omega )a(B,\omega )=\frac{1}{2\pi |\mathrm{\Omega }|}\underset{s=1}{\overset{2}{}}_\mathrm{\Omega }𝒜_{s,s}^{\mathrm{}}(𝐱,𝐱)𝑑𝐱.$$ (3.9) Therefore, the transverse conductivity writes as $$\sigma _{12}(B,\omega )=\frac{e^2}{m^2\omega }a(B,\omega )$$ (3.10) with $`a(B,\omega )`$ given by the r.h.s. of (3.9). A precise formulation of this result is contained in the following theorem: ###### Theorem 3.1. Assume for simplicity that $`\mathrm{\Omega }`$ is the unit cube in $`^3`$. The above defined transverse component of the conductivity tensor admits the thermodynamic limit; more precisely: i. The following operator defined by a $`B(L^2L^2)`$-norm convergent Riemann integral, $`F_L`$ $`:={\displaystyle \frac{1}{2\pi }}{\displaystyle _{\mathrm{\Gamma }_\omega }}f_{FD}(z)\{P_1(B)(H_L(B)z)^1P_2(B)(H_L(B)z\omega )^1`$ $`+zz\omega \}dz,`$ (3.11) is in fact trace-class, and $`\sigma _{12}^{(L)}(B,\omega )=\frac{e^2}{m^2\omega |\mathrm{\Lambda }_L|}\mathrm{Tr}(F_L)`$. ii. Consider the operator $`F_{\mathrm{}}`$ defined by the same integral but with $`H_{\mathrm{}}(B)`$ instead of $`H_L(B)`$, and defined on the whole space. Then $`F_{\mathrm{}}`$ is an integral operator, with a kernel $`𝒜_{s,s^{}}^{\mathrm{}}(𝐱,𝐱^{})`$ jointly continuous on its spatial variables. Moreover, the function defined by $`^3𝐱s_B(𝐱):=_{s=1}^2𝒜_{s,s}^{\mathrm{}}(𝐱,𝐱)`$ is continuous and periodic with respect to $`^3`$. iii. The thermodynamic limit exists: $`\sigma _{12}^{(\mathrm{})}(B,\omega ):=\underset{L\mathrm{}}{lim}\sigma _{12}^{(L)}(B,\omega )={\displaystyle \frac{e^2}{m^2\omega |\mathrm{\Omega }|}}{\displaystyle _\mathrm{\Omega }}s_B(𝐱)𝑑𝐱.`$ (3.12) The proof of this theorem will be given elsewhere . ## 4 The zero frequency limit at $`T=0`$: a rigorous proof of the Widom-Streda formula for semiconductors Doing some very formal computations one can show that at $`T=0`$ and $`\omega =0`$, $`\sigma _{12}(B,\omega )`$ as given by (3.9) and (3.10) coincide with the formula for the quantized Hall conductivity (see e.g. formulas (5),(6) in ) which in turn gives (again at the heuristic level) the well known Widom-Streda formula. The original derivation has little mathematical rigor, in particular because it assumes some very strong assumptions on the existence and regularity of $`(H_{\mathrm{}}(B)\lambda +i0)^1`$ as a function of $`\lambda `$. These assumptions are clearly not true in many situations. Here we will show how the Widom-Streda formula can be rigorously obtained when the Fermi energy lies in a spectral gap. The problem in which the Fermi energy is in the spectrum, remains open. Now assume that for some $`B`$, the chemical potential $`\mu `$ lies in a spectral gap of $`H_{\mathrm{}}(B)`$. More precisely, throughout this section we suppose that $`(d_1,d_2)\rho (H_{\mathrm{}}(B))`$ with $`d_1<d_2`$, and take $`\mu (d_1,d_2)`$. For simplicity, assume that $`\mu =\frac{d_1+d_2}{2}`$. This is the typical situation for semiconductors and/or isolators. In the absence of spin, the Widom-Streda formula roughly states that: $$\sigma _{12}(B,T=0,\omega =0)=ec\frac{N(B,E)}{B}|_{E=\mu },$$ (4.1) where $`N(B,E)`$ is the integrated density of states up to the energy $`E`$. When the spin is present (this was not considered by Streda), this formula is slightly changed. If we denote by $`B_1`$ the $`B`$ multiplying the spin matrix $`\sigma _3`$ in our Hamiltonian (2.1), and with $`B_2`$ the $`B`$ near $`𝐀`$, then in fact we have $$\sigma _{12}(B,T=0,\omega =0)=ec\frac{N(B_1,B_2,E)}{B_2}|_{E=\mu ,B_1=B_2=B}.$$ (4.2) In the rest of this section we give a rigorous (but still not fully technical) proof of (4.2). ###### Theorem 4.1. Consider the conductivity at the thermodynamic limit given in (3.12), and drop the superscript $`\mathrm{}`$. Then if we first take the limit $`T0`$, and after that $`\omega 0`$, we get: $$\underset{\omega 0}{lim}\underset{T0}{lim}\sigma _{12}(B,T,\omega )=ec\frac{}{B_2}\frac{1}{|\mathrm{\Omega }|}\underset{s=1}{\overset{2}{}}_\mathrm{\Omega }\mathrm{\Pi }_{s,s}^B(𝐱,𝐱)𝑑𝐱|_{B_1=B_2=B},$$ (4.3) where $$\mathrm{\Pi }^B=\frac{i}{2\pi }_\mathrm{\Gamma }\frac{1}{H_{\mathrm{}}(B)z}𝑑z$$ (4.4) with a positively oriented contour $`\mathrm{\Gamma }`$ enclosing the spectrum of $`H_{\mathrm{}}(B)`$ below $`\mu `$, i.e. $`\mathrm{\Pi }^B`$ is the Fermi projection onto the subspace of “occupied” states at $`T=0`$. Remarks: 1. Streda did not consider spin in his work and in this case the derivative with respect to the magnetic field appears in the r.h.s. of (4.3). 2. While it is not clear that $`\mathrm{\Pi }^B(𝐱,𝐱)`$ is well defined ($`(H_{\mathrm{}}(B)z)^1(𝐱,𝐱)`$ does not exist!) this can be seen by writing for some $`a\rho (H_{\mathrm{}}(B))`$: $`\mathrm{\Pi }^B`$ $`={\displaystyle \frac{1}{2\pi }}{\displaystyle _\mathrm{\Gamma }}((H_{\mathrm{}}(B)z)^1(H_{\mathrm{}}(B)a)^1dz`$ $`={\displaystyle \frac{1}{2\pi }}{\displaystyle _\mathrm{\Gamma }}(za)(H_{\mathrm{}}(B)z)^1(H_{\mathrm{}}(B)a)^1𝑑z.`$ (4.5) Each resolvent has a polar integral kernel with a $`1/|𝐱𝐱^{}|`$ singularity, and the product of two resolvents will have a continuous kernel. In fact we can repeat this trick and obtain products of as many resolvents as we want, thus further improving the regularity of the integral kernel. Technical details will be given elsewhere. Actually this kind of argument can be used to show that all operators defined by integrals over complex contours have jointly continuous integral kernels. 3. Although the order of limits in (4.3) is important for the argument below, it might be possible (at least under additional conditions on the spectrum of $`H_{\mathrm{}}(B)`$) to interchange the order of limits. The important fact is that the thermodynamic limit has to be taken first: great care is to be taken when defining currents in the static limit for finite systems (for a discussion of this point in a related context see . 4. The result is valid for arbitrary magnetic field $`B`$ and establishes the connection between the Hall conductivity and the Faraday effect. However, the quantum Hall effect requires high magnetic fields while the Faraday effect is usually considered at low magnetic fields. Proof. We start from the conductivity in the thermodynamic limit as given by Theorem 3.1: $$\sigma _{12}(B,T,\omega )=\frac{e^2}{2\pi m^2\omega |\mathrm{\Omega }|}\underset{s=1}{\overset{2}{}}_\mathrm{\Omega }\left[_{\mathrm{\Gamma }_{\beta ,\omega }}f_{FD}(z)\mathrm{\Sigma }(z,\omega )𝑑z\right]_{s,s}(𝐱,𝐱)𝑑𝐱,$$ (4.6) where $$\mathrm{\Sigma }(z,\omega ):=P_1(B)(H_{\mathrm{}}(B)z)^1P_2(B)(H_{\mathrm{}}(B)z\omega )^1+(zz\omega ).$$ (4.7) Since we made the assumption that $`(d_1,d_2)\rho (H_{\mathrm{}}(B)`$, then for $`|\omega |<\frac{d_2d_1}{4}`$ the integral over $`z`$ on the contour $`\mathrm{\Gamma }_{\beta ,\omega }`$ can be replaced with the integral on the contour $`\mathrm{\Gamma }_{\beta ,\omega }^1\mathrm{\Gamma }_{\beta ,\omega }^2`$ where (see also (2.28) and (2.29)): $`\mathrm{\Gamma }_{\beta ,\omega }^1`$ $`=\{x\pm id:axd_1+{\displaystyle \frac{d_2d_1}{4}}\}{\displaystyle \{a+iy:dyd\}}`$ $`{\displaystyle \{d_1+\frac{d_2d_1}{4}+iy:dyd\}}`$ (4.8) and $`\mathrm{\Gamma }_{\beta ,\omega }^2`$ $`=\{x\pm id:xd_2{\displaystyle \frac{d_2d_1}{4}}\}`$ $`{\displaystyle \{d_2\frac{d_2d_1}{4}+iy:dyd\}}.`$ (4.9) Accordingly, one can rewrite $`\sigma _{12}(B,T,\omega )`$ as $`\sigma _{12}(B,T,\omega )={\displaystyle \frac{e^2}{2\pi m^2\omega |\mathrm{\Omega }|}}{\displaystyle \underset{s=1}{\overset{2}{}}}{\displaystyle _\mathrm{\Omega }}`$ (4.10) $`\{{\displaystyle _{\mathrm{\Gamma }_{\beta ,\omega }^1}}\mathrm{\Sigma }(z,\omega )dz`$ $`+{\displaystyle _{\mathrm{\Gamma }_{\beta ,\omega }^1}}(f_{FD}(z)1)\mathrm{\Sigma }(z,\omega )dz+{\displaystyle _{\mathrm{\Gamma }_{\beta ,\omega }^2}}f_{FD}(z)\mathrm{\Sigma }(z,\omega )dz\}_{s,s}(𝐱,𝐱)d𝐱.`$ Note that since the singularities of $`f_{FD}(z)`$ lie on $`\frac{c+d}{2}+iy,y(\mathrm{},\mathrm{})`$, one can take $`\mathrm{\Gamma }_{\beta ,\omega }^j`$ independent of $`\beta `$ i.e. one can take $`d=\frac{|\mathrm{}(\omega )|}{2}`$ in (2.29). At this point we take the limit $`\beta \mathrm{}`$. Since on $`\mathrm{\Gamma }_\omega ^2`$ we have $`|f_{FD}(z)|2\mathrm{exp}[\beta (x\frac{d_1+d_2}{2})]`$, and on $`\mathrm{\Gamma }_\omega ^1`$ we have that $`|f_{FD}(z)1|2\mathrm{exp}[\beta (\frac{d_1+d_2}{2}x)]`$, the last two terms in (4.10) vanish in the zero temperature limit (full details about the control of various integral kernels will be given elsewhere). Hence we get: $`\sigma _{12}(B,T=0,\omega )={\displaystyle \frac{e^2}{2\pi m^2\omega |\mathrm{\Omega }|}}{\displaystyle \underset{s=1}{\overset{2}{}}}{\displaystyle _\mathrm{\Omega }}\left\{{\displaystyle _{\mathrm{\Gamma }_\omega ^1}}\mathrm{\Sigma }(z,\omega )dz\right\}_{s,s}(𝐱,𝐱)d𝐱.`$ (4.11) An application of the Cauchy residue theorem shows that the two terms of $`\mathrm{\Sigma }(z,\omega )`$ (see (4.7)) will combine in the above integral and give $`\sigma _{12}(B,T=0,\omega )={\displaystyle \frac{e^2}{2\pi m^2\omega }}{\displaystyle \frac{1}{|\mathrm{\Omega }|}}{\displaystyle \underset{s=1}{\overset{2}{}}}`$ (4.12) $`{\displaystyle _\mathrm{\Omega }}\left[{\displaystyle _\mathrm{\Gamma }}P_1(B)(H_{\mathrm{}}(B)z+\omega /2)^1P_2(B)(H_{\mathrm{}}(B)z\omega /2)^1\right]_{s,s}(𝐱,𝐱)𝑑𝐱.`$ where $`\mathrm{\Gamma }`$ is any finite contour such that $`\mathrm{\Gamma }\rho (H_{\mathrm{}}(B)+\omega )`$ for all $`|\omega |<\frac{|d_2d_1|}{4}`$ and only enclosing the spectrum of $`H_{\mathrm{}}(B)`$ below $`\frac{d_1+d_2}{2}`$. Now the integrand in (4.12) is analytic in $`\omega `$ in a neighborhood of the origin. By expanding the resolvents one obtains: $`\sigma _{12}(B,T=0,\omega )=`$ (4.13) $`{\displaystyle \frac{e^2}{2\pi m^2}}{\displaystyle \frac{1}{|\mathrm{\Omega }|}}{\displaystyle \underset{s=1}{\overset{2}{}}}{\displaystyle _\mathrm{\Omega }}\{{\displaystyle _\mathrm{\Gamma }}{\displaystyle \frac{1}{\omega }}P_1(B)(H_{\mathrm{}}(B)z)^1P_2(B)(H_{\mathrm{}}(B)z)^1`$ $`+{\displaystyle \frac{1}{2}}[P_1(B)(H_{\mathrm{}}(B)z)^1P_2(B)(H_{\mathrm{}}(B)z)^2`$ $`P_1(B)(H_{\mathrm{}}(B)z)^2P_2(B)(H_{\mathrm{}}(B)z)^1]{\displaystyle \frac{}{}}\}_{s,s}(𝐱,𝐱)d𝐱+𝒪(\omega ).`$ Apparently we have a first oder pole at $`\omega =0`$. But we now prove that the singular term in the r.h.s. of (4.13) is identically zero. Namely (when no spin variables appear the integral kernels below have to be understood as matrices in the spin space): $$\left\{_\mathrm{\Gamma }P_1(B)(H_{\mathrm{}}(B)z)^1P_2(B)(H_{\mathrm{}}(B)z)^1𝑑z\right\}(𝐱,𝐱)=0.$$ (4.14) Using the magnetic perturbation theory and the trick from (2) one can prove that even though the integrand in (4.14) has a quite singular kernel, after integration with respect to $`z`$ one gets a smooth kernel, exponentially localized near the diagonal (details will be given elsewhere). Let us notice an operator equality which makes sense on compactly supported functions: $`{\displaystyle \frac{1}{m}}(H_{\mathrm{}}(B)z)^1P_2(B)(H_{\mathrm{}}(B)z)^1=i[X_2,(H_{\mathrm{}}(B)z)^1],`$ (4.15) because the resolvent $`(H_{\mathrm{}}(B)z)^1`$ sends compactly supported functions into exponentially decaying functions (see (3.7)), which are in the domain of $`X_2`$. In fact, the operator on the right side has a nice integral kernel, given by $`\left\{i[X_2,(H_{\mathrm{}}(B)z)^1]\right\}(𝐱,𝐲)=i(x_2y_2)G_{\mathrm{}}(𝐱,𝐲;z),`$ (4.16) which is no longer singular at the diagonal and still exponentially localized near the diagonal, thus defining a bounded operator on the whole Hilbert space. After integration we get: $`{\displaystyle \frac{i}{2\pi }}{\displaystyle _\mathrm{\Gamma }}P_1(B)(H_{\mathrm{}}(B)z)^1P_2(B)(H_{\mathrm{}}(B)z)^1𝑑z=`$ (4.17) $`{\displaystyle \frac{im}{2\pi }}{\displaystyle _\mathrm{\Gamma }}P_1(B)[(H_{\mathrm{}}(B)z)^1,X_2]𝑑z=m[P_1(B)\mathrm{\Pi }^B,X_2],`$ where we used that $`P_1(B)`$ and $`X_2`$ commute. Note that the magnetic perturbation theory states that the integral kernel of $`P_1(B)\mathrm{\Pi }^B`$ is smooth and exponentially localized near the diagonal. Therefore $`[P_1(B)\mathrm{\Pi }^B,X_2]`$ will have the integral kernel: $$\left\{[P_1(B)\mathrm{\Pi }^B,X_2]\right\}(𝐱,𝐲)=(y_2x_2)\{P_1(B)\mathrm{\Pi }^B\}(𝐱,𝐲)$$ (4.18) which is identically zero at the diagonal and proves (4.14). We can conclude at this point that: $`\underset{\omega 0}{lim}\sigma _{12}(B,T=0,\omega )=`$ (4.19) $`{\displaystyle \frac{e^2}{4\pi m^2}}{\displaystyle \frac{1}{|\mathrm{\Omega }|}}{\displaystyle \underset{s=1}{\overset{2}{}}}{\displaystyle _\mathrm{\Omega }}\{{\displaystyle \frac{}{}}P_1(B)(H_{\mathrm{}}(B)z)^1P_2(B)(H_{\mathrm{}}(B)z)^2`$ $`P_1(B)(H_{\mathrm{}}(B)z)^2P_2(B)(H_{\mathrm{}}(B)z)^1{\displaystyle \frac{}{}}\}_{s,s}(𝐱,𝐱)d𝐱.`$ Now consider the r.h.s. of (4.3). Since the magnetic field multiplying the spin will not change, our notation will only refer to $`B_2`$. Due to the stability of the spectrum against small variations of the magnetic field, for sufficiently small $`\mathrm{\Delta }B`$, $`\mathrm{\Pi }^{B_2+\mathrm{\Delta }B}`$ still exists and $$\mathrm{\Pi }^{B_2+\mathrm{\Delta }B}\mathrm{\Pi }^{B_2}=\frac{i}{2\pi }_\mathrm{\Gamma }((H_{\mathrm{}}(B_2+\mathrm{\Delta }B)z)^1(H_{\mathrm{}}(B_2)z)^1)𝑑z.$$ (4.20) By using the magnetic perturbation theory with respect to $`\mathrm{\Delta }B`$ (see also the discussion around (6.2)) one obtains: $`[\mathrm{\Pi }^{B_2+\mathrm{\Delta }B}\mathrm{\Pi }^{B_2}](𝐱,𝐱)={\displaystyle \frac{ie\mathrm{\Delta }B}{2\pi mc}}{\displaystyle _\mathrm{\Gamma }}\{{\displaystyle _^3}d𝐲(H_{\mathrm{}}(B_2)z)^1(𝐱,𝐲)`$ (4.21) $`{\displaystyle \frac{}{}}[𝐏_𝐲(B_2)𝐀(𝐲𝐱)](H_{\mathrm{}}(B_2)z)^1(𝐲,𝐱)\}dz+𝒪((\mathrm{\Delta }B)^2).`$ Now a very important identity is (see (3.3), (4.15) and (4.16)) $`𝐀(𝐲𝐱)(H_{\mathrm{}}(B_2)z)^1(𝐲,𝐱)=`$ (4.22) $`{\displaystyle \frac{i}{2m}}𝐧_3[(H_{\mathrm{}}(B_2)z)^1𝐏(B_2)(H_{\mathrm{}}(B_2)z)^1](𝐲,𝐱).`$ The remainder in $`(\mathrm{\Delta }B)^2`$ will have a smooth integral kernel after the integration with respect to $`z`$, hence we obtain: $`{\displaystyle \frac{}{B_2}}{\displaystyle \frac{1}{|\mathrm{\Omega }|}}{\displaystyle _\mathrm{\Omega }}\mathrm{\Pi }^B(𝐱,𝐱)𝑑𝐱={\displaystyle \frac{e}{4\pi m^2c}}`$ (4.23) $`{\displaystyle _\mathrm{\Omega }}\{{\displaystyle _\mathrm{\Gamma }}[{\displaystyle \frac{}{}}(H_{\mathrm{}}(B)z)^1P_1(B)(H_{\mathrm{}}(B)z)^1P_2(B)(H_{\mathrm{}}(B)z)^1`$ $`{\displaystyle \frac{}{}}(H_{\mathrm{}}(B)z)^1P_2(B)(H_{\mathrm{}}(B)z)^1P_1(B)(H_{\mathrm{}}(B)z)^1)dz]\}(𝐱,𝐱).`$ From (4.23) and (4.19) we see that (4.3) follows if we can prove that one can circularly permute the operators under the integral sign in (4.23). One can prove this by interpreting (4.23) as the thermodynamic limit of the corresponding expression on finite volume and then using the invariance of the trace under cyclic permutations. Alternatively one can prove it directly and in what follows we outline the proof. Due to the smoothing effect of the integral with respect to $`z`$, we can always restrict ourselves to considering a product of only two integral operators which commute with the discrete magnetic translations, and have kernels $`e^{ib\varphi _𝐚(𝐱,𝐲)}K_1(𝐱,𝐲)`$ and $`e^{ib\varphi _𝐚(𝐲,𝐱^{})}K_2(𝐲,𝐱^{})`$. We therefore look at an absolutely convergent integral of the form (the anti-symmetric magnetic phases disappear when we look at the diagonal, see (3.1)): $$_\mathrm{\Omega }𝑑𝐱_^3𝑑𝐲K_1(𝐱,𝐲)K_2(𝐲,𝐱)$$ (4.24) with $$K_{1,2}(𝐱,𝐲)=K_{1,2}(𝐱+\stackrel{}{\gamma },𝐲+\stackrel{}{\gamma }).$$ Then $`{\displaystyle _\mathrm{\Omega }}𝑑𝐱{\displaystyle _^3}𝑑𝐲K_1(𝐱,𝐲)K_2(𝐲,𝐱)={\displaystyle \underset{\stackrel{}{\gamma }}{}}{\displaystyle _\mathrm{\Omega }}𝑑𝐱{\displaystyle _\mathrm{\Omega }}𝑑𝐲K_1(𝐱,𝐲+\stackrel{}{\gamma })K_2(𝐲+\stackrel{}{\gamma },𝐱)`$ $`={\displaystyle \underset{\stackrel{}{\gamma }}{}}{\displaystyle _\mathrm{\Omega }}𝑑𝐱{\displaystyle _\mathrm{\Omega }}𝑑𝐲K_1(𝐱\stackrel{}{\gamma },𝐲)K_2(𝐲,𝐱\stackrel{}{\gamma })`$ $`={\displaystyle _\mathrm{\Omega }}𝑑𝐲{\displaystyle _^3}𝑑𝐱{\displaystyle _^3}K_2(𝐲,𝐱)K_1(𝐱,𝐲)`$ (4.25) which gives the needed “trace cyclicity” and the theorem is proved.∎ We now turn to the question whether the limit in (4.3) actually vanishes as is suggested by some heuristic arguments (see e.g. ). We start by recalling some results about Wannier functions. Let $`\sigma _0(B_0)`$ be an isolated part of the spectrum of $`H_{\mathrm{}}(B_0)`$ and $`\mathrm{\Pi }_0^{B_0}`$ the corresponding spectral projection. We say that $`\mathrm{\Pi }_0^{B_0}`$ has a basis of exponentially localized (magnetic) Wannier functions if there exist $`\alpha >0`$, $`w_jL^2(^3)L^2(^3),j=1,2,\mathrm{},p<\mathrm{}`$ satisfying (we denote by $`w_j(𝐱,s)`$ the values of $`w_j`$ in $`𝐱^3`$ and $`s\{1,2\}`$) $$\underset{s=1}{\overset{2}{}}_^3|w_j(𝐱,s)|^2e^{2\alpha |𝐱|}𝑑𝐱M<\mathrm{},$$ (4.26) such that the set of functions $`\left\{w_{j,\stackrel{}{\gamma }}\right\}_{j=1,2,\mathrm{}p,\stackrel{}{\gamma }}`$ with $$w_{j,\stackrel{}{\gamma }}(𝐱,s)=e^{ib\varphi _𝐚(𝐱,\stackrel{}{\gamma })}w_j(𝐱\stackrel{}{\gamma },s)$$ is a basis in the range of the projection $`\mathrm{\Pi }_0^{B_0}(L^2(^3)L^2(^3))`$. If the spin is neglected, it has been proved in that the existence of bases of exponentially localized Wannier functions is stable against small values of the magnetic field (i.e. $`B_0=0`$). More precisely, if $`\sigma _0`$ is an isolated part of the spectrum of $`\mathrm{\Delta }+V`$ and the corresponding subspace has a basis of exponentially localized Wannier functions then, for sufficiently small $`B`$, $`\sigma _0(B)`$ is still isolated and the corresponding spectral subspace has a basis of exponentially localized magnetic Wannier functions. The methods in together with the magnetic perturbation theory , , allow one to generalize the above result to arbitrary $`B_0`$ and presence of the spin (as far as the spin-orbit term is sufficiently small) . Now the existence of exponentially localized magnetic Wannier functions for an isolated part of the spectrum and for the value of the magnetic field $`B_2`$ in an interval around $`B_0`$ allows one to write: $`{\displaystyle \underset{s=1}{\overset{2}{}}}{\displaystyle _\mathrm{\Omega }}\mathrm{\Pi }_{0;s,s}^B(𝐱,𝐱)𝑑𝐱={\displaystyle _\mathrm{\Omega }}{\displaystyle \underset{j,\stackrel{}{\gamma }}{}}{\displaystyle \underset{s=1}{\overset{2}{}}}|w_{j,\stackrel{}{\gamma }}(𝐱,s)|^2d𝐱`$ $`={\displaystyle \underset{j,\stackrel{}{\gamma }}{}}{\displaystyle \underset{s=1}{\overset{2}{}}}{\displaystyle _\mathrm{\Omega }}𝑑𝐱|w_j(𝐱\stackrel{}{\gamma },s)|^2={\displaystyle \underset{j=1}{\overset{p}{}}}{\displaystyle \underset{s=1}{\overset{2}{}}}{\displaystyle _^3}𝑑𝐱|w_j(𝐱,s)|^2=p.`$ (4.27) Thus the integrated density of states corresponding to the Fermi projection is constant in $`B_2`$ in a small interval around $`B_0`$, hence this band gives no contribution in the r.h.s. of (4.3). For small fields, the above discussion can be summarized in: ###### Theorem 4.2. Suppose $`(d_1,d_2)\rho (\mathrm{\Delta }+V)`$, $`d_2>d_1`$, and that the spectral subspace corresponding to $`(\mathrm{},d_1]`$ admits a basis of exponentially localized Wannier functions. Suppose that the spin-orbit interaction (see (2.2)) is small enough such that as $`c^2`$ decreases from $`\mathrm{}`$ to its actual value, we have that $`\frac{d_1+d_2}{2}\rho (H_{\mathrm{}}(0))`$. Then for sufficiently small $`B`$: $$\underset{\omega 0}{lim}\underset{T0}{lim}\sigma _{12}(B,T,\omega )=0.$$ (4.28) In particular all the derivatives of $`\sigma _{12}(B,0)`$ vanish for $`B=0`$, and this substantiates Roth’s result (formula (50) in ) for the first order correction in $`B`$ at zero frequency. ## 5 A closed formula for free electrons If $`V=0`$ it turns out that the conductivity tensor can be explicitly computed for all values of $`B`$ and $`\omega `$. The formula does not depend on whether we work in two or three dimensions. More precisely, we will show in this section that $$\sigma _{12}(B,\omega )=\frac{e^3n}{m^2c}\frac{B}{\omega ^2\frac{B^2e^2}{m^2c^2}},$$ (5.1) where $`n=n(T,\mu ,B)`$ is the grand-canonical density. The formula (5.1) is well known in classical physics and goes back at least to Drude but we are not aware of a known fully quantum derivation. The coincidence of classical and quantum formulas can be understood taking into account that the Hamiltonians involved (choose the symmetric gauge) are quadratic and it is known that for this class of operators classical and quantum computations coincide in many instances. While it is possible to derive (5.1) by using the explicit form of the Green function or alternatively of eigenvalues and eigenprojections for the Landau Hamiltonian (see e.g. ) we shall obtain it below only using resolvent and commutation identities. Let us only notice that when $`\omega =0`$ we re-obtain formula (18) in , while for a fixed frequency we get $$\frac{\sigma _{12}}{B}(0,\omega )=\frac{e^3n}{m^2c\omega ^2}$$ which is “the high frequency limit” or what Roth also calls “the free electron Faraday effect” in formula (51) from . We begin by listing a few identities which are valid for a free electron on the entire space. $`i[P_1(B),P_2(B)]={\displaystyle \frac{Be}{c}}`$ $`i[H_{\mathrm{}}(B),P_1(B)]={\displaystyle \frac{Be}{mc}}P_2(B),`$ $`i[H_{\mathrm{}}(B),P_2(B)]={\displaystyle \frac{Be}{mc}}P_1(B),`$ $`[H_{\mathrm{}}(B),[H_{\mathrm{}}(B),P_1(B)]]={\displaystyle \frac{B^2e^2}{m^2c^2}}P_1(B),`$ (5.2) $`[H_{\mathrm{}}(B),[H_{\mathrm{}}(B),P_2(B)]]={\displaystyle \frac{B^2e^2}{m^2c^2}}P_2(B).`$ Next, since in this case $`𝒜_{s,s}^{\mathrm{}}(𝐱,𝐱)`$ does not depend upon $`𝐱`$ one has $`a(B,\omega )={\displaystyle \frac{1}{2\pi }}{\displaystyle \underset{s=1}{\overset{2}{}}}`$ (5.3) $`\{{\displaystyle _{\mathrm{\Gamma }_\omega }}dzf_{FD}(z)[P_1(B)(H_{\mathrm{}}(B)z)^1P_2(B)(H_{\mathrm{}}(B)z\omega )^1`$ $`+{\displaystyle \frac{}{}}zz\omega ]\}(\stackrel{}{0},s;\stackrel{}{0},s).`$ Commuting $`(H_{\mathrm{}}(B)z\omega )^1`$ with $`P_2(B)`$ in the first term, and $`P_1(B)`$ with $`(H_{\mathrm{}}(B)z+\omega )^1`$ in the second one, we obtain $`a(B,\omega )={\displaystyle \frac{1}{2\pi |\mathrm{\Omega }|}}{\displaystyle \underset{s=1}{\overset{2}{}}}`$ $`\{{\displaystyle _{\mathrm{\Gamma }_\omega }}dzf_{FD}(z)[P_1(B)(H_{\mathrm{}}(B)z)^1(H_{\mathrm{}}(B)z\omega )^1P_2(B){\displaystyle \frac{}{}}`$ $`+P_1(B)(H_{\mathrm{}}(B)z)^1(H_{\mathrm{}}(B)z\omega )^1`$ $`[H_{\mathrm{}}(B),P_2(B)](H_{\mathrm{}}(B)z\omega )^1`$ $`+(H_{\mathrm{}}(B)z+\omega )^1P_1(B)P_2(B)(H_{\mathrm{}}(B)z)^1`$ $`+(H_{\mathrm{}}(B)z+\omega )^1[H_{\mathrm{}}(B),P_1(B)]`$ $`{\displaystyle \frac{}{}}(H_{\mathrm{}}(B)z+\omega )^1P_2(B)(H_{\mathrm{}}(B)z)^1]\}(\stackrel{}{0},s;\stackrel{}{0},s)`$ $`=I+II+III+IV.`$ (5.4) Now $`I+III`$ can easily be computed. Indeed, by cyclic permutations (see (4)) one can cluster the two resolvents and then by the resolvent identity $$(Az_1)^1(Az_2)^1=(z_1z_2)^1[(Az_1)^1(Az_2)^1],$$ (5.5) one obtains four terms. Two of them vanish after the integration over $`z`$ due to the analyticity of the integrand while the other two give $`I+III={\displaystyle \frac{1}{2\pi }}{\displaystyle \underset{s=1}{\overset{2}{}}}`$ (5.6) $`\left\{{\displaystyle _{\mathrm{\Gamma }_\omega }}𝑑zf_{FD}(z){\displaystyle \frac{1}{\omega }}[P_2(B),P_1(B)](H_{\mathrm{}}(B)z)^1\right\}(\stackrel{}{0},s;\stackrel{}{0},s)`$ $`={\displaystyle \frac{Be}{\omega }}{\displaystyle \underset{s=1}{\overset{2}{}}}\{f_{FD}(H_{\mathrm{}}(B))\}(\stackrel{}{0},s;\stackrel{}{0},s)=:{\displaystyle \frac{Be}{\omega }}n(T,\mu ,B).`$ (5.7) In an analogous manner $`III+IV={\displaystyle \frac{1}{2\pi \omega }}{\displaystyle \underset{s=1}{\overset{2}{}}}\{{\displaystyle _{\mathrm{\Gamma }_\omega }}dzf_{FD}(z)`$ (5.8) $`\{(H_{\mathrm{}}(B)z)^1[H_{\mathrm{}}(B),P_2(B)](H_{\mathrm{}}(B)z\omega )^1P_1(B)`$ $`(H_{\mathrm{}}(B)z)^1[H_{\mathrm{}}(B),P_1(B)](H_{\mathrm{}}(B)z+\omega )^1P_2(B)\}`$ $`{\displaystyle \frac{}{}}\}(\stackrel{}{0},s;\stackrel{}{0},s).`$ At this point we commute $`[H_{\mathrm{}}(B),P_2(B)]`$ with $`(H_{\mathrm{}}(B)z\omega )^1`$ in the first term, and $`[H_{\mathrm{}}(B),P_1(B)]`$ with $`(H_{\mathrm{}}(B)z+\omega )^1`$ in the second term and use (3.3) again. Some of the terms vanish after performing the integration over $`z`$ and the remaining ones write as: $`{\displaystyle \frac{1}{\omega }}(H_{\mathrm{}}(B)z)^1[H_{\mathrm{}}(B),P_2(B)]P_1(B)`$ (5.9) $`{\displaystyle \frac{1}{\omega }}(H_{\mathrm{}}(B)z)^1[H_{\mathrm{}}(B),[H_{\mathrm{}}(B),P_2(B)]](H_{\mathrm{}}(B)z\omega )^1P_1(B)`$ $`{\displaystyle \frac{1}{\omega }}(H_{\mathrm{}}(B)z)^1[H_{\mathrm{}}(B),P_1(B)]P_2(B)`$ $`{\displaystyle \frac{1}{\omega }}(H_{\mathrm{}}(B)z)^1[H_{\mathrm{}}(B),[H_{\mathrm{}}(B),P_1(B)]](H_{\mathrm{}}(B)z+\omega )^1P_2(B).`$ Taking into account (5) the first and the third terms in (5.9) combine to $$\frac{1}{\omega }(H_{\mathrm{}}(B)z)^1[H_{\mathrm{}}(B),P_1(B)P_2(B)]$$ which after integration over $`z`$ is proportional to $`f_{FD}(H_{\mathrm{}}(B))i[H_{\mathrm{}}(B),P_1(B)P_2(B)]`$ (5.10) $`={\displaystyle \frac{Be}{mc}}\left\{f_{FD}(H_{\mathrm{}}(B))P_1(B)^2f_{FD}(H_{\mathrm{}}(B))P_2(B)^2\right\},`$ where we used the second and third identities in (5). Consider the unitary operator $`U`$ which implements the coordinate change $`(Uf)(x_1,x_2,x_3)=f(x_2,x_1,x_3)`$. Then one can prove that $`UP_1(B)U^{}=P_2(B)`$, $`UP_2(B)U^{}=P_1(B)`$ and $`UH_{\mathrm{}}(B)U^{}=H_{\mathrm{}}(B)`$. This implies that $$Uf_{FD}(H_{\mathrm{}}(B))P_1(B)^2U^{}=f_{FD}(H_{\mathrm{}}(B))P_2(B)^2.$$ Since both operators have a smooth integral kernel, and because the rotation with $`U`$ does not change the diagonal value of the integral kernel on the left hand side, it means that the contribution given by (5.10) is zero. Therefore we only remain with the second and fourth terms in (5.9). Using (5), they become: $`{\displaystyle \frac{B^2e^2}{m^2c^2\omega }}(H_{\mathrm{}}(B)z)^1P_2(B)(H_{\mathrm{}}(B)z\omega )^1P_1(B)`$ $`{\displaystyle \frac{B^2e^2}{m^2c^2\omega }}(H_{\mathrm{}}(B)z)^1P_1(B)(H_{\mathrm{}}(B)z+\omega )^1P_2(B).`$ (5.11) Using once more the cyclicity of the trace and comparing with the starting point (5.3), we obtain the remarkable identity $$II+IV=\frac{B^2e^2}{m^2c^2\omega ^2}a(B,\omega ).$$ (5.12) Putting together (5), (5.6), and (5.12), we obtain the equation: $$a(B,\omega )=\frac{Be}{c\omega }n+\frac{B^2e^2}{m^2c^2\omega ^2}a(B,\omega ),$$ which gives (5.1) (see (3.10)). ## 6 Magnetic perturbation theory and the linear term in $`B`$ When $`V0`$ it is no longer possible to obtain a closed formula for $`\sigma _{12}(B,\omega )`$. Since in most physical applications the external magnetic field can be considered weak, an expansion in $`B`$ up to the first or second order would be sufficient. In this section we show that $`a_L(B,\omega )`$ has an expansion in $`B`$ to any order and write down the expressions of the first two terms. The first one gives the transverse conductivity at zero magnetic field and the second which is linear in $`B`$ provides the Verdet constant. From (3.6) and (3.8) (in what follows by $`tr`$ we mean the trace over the spin variable): $`a_L(B,\omega )={\displaystyle \frac{1}{2\pi |\mathrm{\Lambda }_L|}}{\displaystyle _{\mathrm{\Lambda }_L}}d𝐱\{\mathrm{tr}{\displaystyle _{\mathrm{\Gamma }_\omega }}dzf_{FD}(z)`$ (6.1) $`{\displaystyle _{\mathrm{\Lambda }_L}}d𝐮\{\left[(P_{𝐱,1}(0)bA_1(𝐱𝐮))K_L(𝐱,𝐮;z)\right]`$ $`\left[(P_{𝐮,2}(0)bA_2(𝐮𝐱^{}))K_L(𝐮,𝐱^{};z+\omega )\right]`$ $`+\left[(P_{𝐱,1}(0)bA_1(𝐱𝐮))K_L(𝐱,𝐮;z\omega )\right]`$ $`{\displaystyle \frac{}{}}\left[(P_{𝐮,2}(0)bA_2(𝐮𝐱^{}))K_L(𝐮,𝐱^{};z)\right]\}\left\}\right|_{𝐱=𝐱^{}}`$ Let us mention here that one cannot interchange the order of the above integrals. First one performs the integral with respect to $`𝐮`$, then the integral in $`z`$, then we can put $`𝐱=𝐱^{}`$ since the resulting kernel is smooth, and finally one integrates with respect to $`𝐱`$ over $`\mathrm{\Lambda }_L`$. When considering the expansion in $`b`$ of $`a_L(B,\omega )`$ we are left with the problem of the expansion of $`K_{\mathrm{\Lambda }_L}(𝐱,𝐲;\zeta )`$ . This expansion is provided by the magnetic perturbation theory as developed in . Following the steps in in the case at hand one obtains: $`K_L(𝐱,𝐲;z)=G_L^{(0)}(𝐱,𝐲;z)`$ (6.2) $`+{\displaystyle \frac{b}{m}}{\displaystyle _{\mathrm{\Lambda }_L}}G_L^{(0)}(𝐱,𝐮;z)\left[𝐏_𝐮(0)𝐀(𝐮𝐲)G_L^{(0)}(𝐮,𝐲;z)\right]𝑑𝐮`$ $`+b{\displaystyle \frac{gc\mu _b}{e}}{\displaystyle _{\mathrm{\Lambda }_L}}G_L^{(0)}(𝐱,𝐮;z)\sigma _3G_L^{(0)}(𝐮,𝐲;z)𝑑𝐮+𝒪(b^2)`$ $`=G_L^{(0)}(𝐱,𝐲;z)+bG_L^{(orbit)}(𝐱,𝐲;z)+bG_L^{(spin)}(𝐱,𝐲;z)+𝒪(b^2).`$ The above integrands are matrices in the spin variable, that is why the spin does not appear explicitly. The error term $`𝒪(b^2)`$ can also be fully controlled with the magnetic perturbation theory (actually arbitrary order terms can be computed; see for details). Plugging the expansion (6.2) into (6.2) and collecting the terms of zero and first order one obtains $`a_L(B,\omega )=a_L(0,\omega )+ba_{L,1}(\omega )+𝒪(b^2),`$ (6.3) where the zeroth order term is: $`a_L(0,\omega )={\displaystyle \frac{1}{2\pi |\mathrm{\Lambda }_L|}}{\displaystyle _{\mathrm{\Lambda }_L}}d𝐱\{\mathrm{tr}{\displaystyle _{\mathrm{\Gamma }_\omega }}dzf_{FD}(z)`$ (6.4) $`{\displaystyle \frac{}{}}\{P_1(0)(H_L(0)z)^1P_2(0)(H_L(0)z+\omega )^1+(zz\omega )\}\left\}\right|_{𝐱=𝐱^{}},`$ while the first order correction reads as: $$a_{L,1}(\omega )=a_{L,1}^{orbit}(\omega )+a_{L,1}^{spin}(\omega ),$$ (6.5) where $`a_{L,1}^{orbit}(\omega )={\displaystyle \frac{1}{2\pi |\mathrm{\Lambda }_L|}}{\displaystyle _{\mathrm{\Lambda }_L}}d𝐱\{\mathrm{tr}{\displaystyle _{\mathrm{\Gamma }_\omega }}dzf_{FD}(z)`$ (6.6) $`{\displaystyle _{\mathrm{\Lambda }_L}}d𝐮\{\left[A_1(𝐱𝐮)G_L^{(0)}(𝐱,𝐮;z)\right]\left[P_{𝐮,2}(0)G_L^{(0)}(𝐮,𝐱^{};z+\omega )\right]`$ $`\left[P_{𝐱,1}(0)G_L^{(0)}(𝐱,𝐮;z)\right]\left[A_2(𝐮𝐱^{})G_L^{(0)}(𝐮,𝐱^{};z+\omega )\right]`$ $`+\left[P_{𝐱,1}(0)G_L^{(orbit)}(𝐱,𝐮;z)\right]\left[P_{𝐮,2}(0)G_L^{(0)}(𝐮,𝐱^{};z+\omega )\right]`$ $`+\left[P_{𝐱,1}(0)G_L^{(0)}(𝐱,𝐮;z)\right]\left[P_{𝐮,2}(0)G_L^{(orbit)}(𝐮,𝐱^{};z+\omega )\right]`$ $`+{\displaystyle \frac{}{}}(zz\omega )\}\left\}\right|_{𝐱=𝐱^{}},`$ $`a_{L,1}^{spin}(\omega )={\displaystyle \frac{1}{2\pi |\mathrm{\Lambda }_L|}}{\displaystyle _{\mathrm{\Lambda }_L}}d𝐱\{\mathrm{tr}{\displaystyle _{\mathrm{\Gamma }_\omega }}dzf_{FD}(z)`$ (6.7) $`{\displaystyle _{\mathrm{\Lambda }_L}}d𝐮\{\left[P_{𝐱,1}(0)G_L^{(spin)}(𝐱,𝐮;z)\right]\left[P_{𝐮,2}(0)G_L^{(0)}(𝐮,𝐱^{};z+\omega )\right]`$ $`+\left[P_{𝐱,1}(0)G_L^{(0)}(𝐱,𝐮;z)\right]\left[P_{𝐮,2}(0)G_L^{(spin)}(𝐮,𝐱^{};z+\omega )\right]`$ $`+{\displaystyle \frac{}{}}(zz\omega )\}\left\}\right|_{𝐱=𝐱^{}}.`$ Now consider the expression $`𝐀(𝐱𝐲)G_L^{(0)}(𝐱,𝐲;z)`$ appearing in the formula for $`a_{L,1}(\omega )`$. Observing that it represents a commutator (see (3.3)) one has the identity $`𝐀(𝐱𝐲)G_L^{(0)}(𝐱,𝐲;z)=\left({\displaystyle \frac{1}{2}}𝐧_3(𝐱𝐲)\right)G_L^{(0)}(𝐱,𝐲;z)`$ $`=\left({\displaystyle \frac{1}{2}}𝐧_3[𝐗,(H_L(0)z)^1]\right)(𝐱,𝐲)`$ $`={\displaystyle \frac{i}{2m}}\{(H_L(0)z)^1(𝐧_3P)(H_L(0)z)^1\}(𝐱,𝐲),`$ (6.8) where $`𝐗`$ denotes the multiplication operator with $`𝐱`$. By a straightforward (but somewhat tedious) computation one arrives at: $$a_{L,1}(\omega )=a_{L,1}^{orbit,1}(\omega )+a_{L,1}^{orbit,2}(\omega )+a_{L,1}^{spin}(\omega )$$ (6.9) where $`a_{L,1}^{orbit,1}(\omega )`$ $`={\displaystyle \frac{i}{4m\pi \omega |\mathrm{\Lambda }_L|}}{\displaystyle _{\mathrm{\Lambda }_L}}d𝐱\{\mathrm{tr}{\displaystyle _{\mathrm{\Gamma }_\omega }}dzf_{FD}(z)`$ (6.10) $`[{\displaystyle \frac{}{}}{\displaystyle \underset{\alpha =1}{\overset{2}{}}}P_\alpha (0)(H_L(0)z)^1P_\alpha (0)(H_L(0)z\omega )^1`$ $`+{\displaystyle \underset{\alpha =1}{\overset{2}{}}}P_\alpha (0)(H_L(0)z)^1P_\alpha (0)(H_L(0)z+\omega )^1`$ $`{\displaystyle \frac{}{}}{\displaystyle \underset{\alpha =1}{\overset{2}{}}}P_\alpha (0)(H_L(0)z)^1P_\alpha (0)(H_L(0)z)^1]\}(𝐱,𝐱),`$ $`a_{L,1}^{orbit,2}(\omega )`$ $`={\displaystyle \frac{i}{4\pi m^2|\mathrm{\Lambda }_L|}}{\displaystyle _{\mathrm{\Lambda }_L}}d𝐱\{\mathrm{tr}{\displaystyle _{\mathrm{\Gamma }_\omega }}dzf_{FD}(z)`$ (6.11) $`\{P_1(0)(H_L(0)z)^1P_1(0)(H_L(0)z)^1`$ $`P_2(0)(H_L(0)z)^1P_2(0)(H_L(0)z\omega )^1`$ $`+P_1(0)(H_L(0)z)^1P_2(0)(H_L(0)z)^1`$ $`P_1(0)(H_L(0)z)^1P_2(0)(H_L(0)z\omega )^1`$ $`P_1(0)(H_L(0)z+\omega )^1P_1(0)(H_L(0)z+\omega )^1`$ $`P_2(0)(H_L(0)z+\omega )^1P_2(0)(H_L(0)z)^1`$ $`+P_1(0)(H_L(0)z+\omega )^1P_2(0)(H_L(0)z+\omega )^1`$ $`P_1(0)(H_L(0)z+\omega )^1P_2(0)(H_L(0)z)^1`$ $`P_1(0)(H_L(0)z)^1P_2(0)(H_L(0)z\omega )^1`$ $`P_1(0)(H_L(0)z\omega )^1P_2(0)(H_L(0)z\omega )^1`$ $`+P_1(0)(H_L(0)z)^1P_2(0)(H_L(0)z\omega )^1`$ $`P_2(0)(H_L(0)z\omega )^1P_1(0)(H_L(0)z\omega )^1`$ $`P_1(0)(H_L(0)z+\omega )^1P_2(0)(H_L(0)z)^1`$ $`P_1(0)(H_L(0)z)^1P_2(0)(H_L(0)z)^1`$ $`+P_1(0)(H_L(0)z+\omega )^1P_2(0)(H_L(0)z)^1`$ $`{\displaystyle \frac{}{}}P_2(0)(H_L(0)z)^1P_1(0)(H_L(0)z)^1\}\}(𝐱,𝐱),`$ and $`a_{L,1}^{spin}(\omega )={\displaystyle \frac{gc\mu _b}{2e\pi |\mathrm{\Lambda }_L|}}{\displaystyle _{\mathrm{\Lambda }_L}}d𝐱\{\mathrm{tr}{\displaystyle _{\mathrm{\Gamma }_\omega }}dzf_{FD}(z)`$ (6.12) $`\{\left[P_1(0)(H_L(0)z)^1\sigma _3(H_L(0)z)^1P_2(0)(H_L(0)z\omega )^1\right]`$ $`+\left[P_1(0)(H_L(0)z)^1P_2(0)(H_L(0)z\omega )^1\sigma _3(H_L(0)z\omega )^1\right]`$ $`+\left[P_1(0)(H_L(0)z+\omega )^1\sigma _3(H_L(0)z+\omega )^1P_2(0)(H_L(0)z)^1\right]`$ $`{\displaystyle \frac{}{}}+\left[P_1(0)(H_L(0)z+\omega )^1P_2(0)(H_L(0)z)^1\sigma _3(H_L(0)z)^1\right]\}\}(𝐱,𝐱).`$ ## 7 The periodic case Now consider the case when $`V`$ is periodic. In this case, after taking the thermodynamic limit one can replace (see (3.9)) $`\frac{1}{|\mathrm{\Lambda }_L|}_{\mathrm{\Lambda }_L}`$ with $`\frac{1}{|\mathrm{\Omega }|}_\mathrm{\Omega }`$ and rewrite (6.10)-(6.12) as integrals over the Brillouin zone $`a_{\mathrm{},1}^{orbit,1}(\omega )`$ $`={\displaystyle \frac{i}{4m\pi \omega |\mathrm{\Omega }|}}{\displaystyle _\mathrm{\Omega }}d𝐤{\displaystyle _\mathrm{\Omega }}d𝐱\{\mathrm{tr}{\displaystyle _{\mathrm{\Gamma }_\omega }}dzf_{FD}(z)`$ (7.1) $`{\displaystyle \underset{\alpha =1}{\overset{2}{}}}(p_\alpha +k_\alpha )(h(𝐤)z)^1(p_\alpha +k_\alpha )(h(𝐤)z\omega )^1`$ $`+{\displaystyle \underset{\alpha =1}{\overset{2}{}}}(p_\alpha +k_\alpha )(h(𝐤)z)^1(p_\alpha +k_\alpha )(h(𝐤)z+\omega )^1`$ $`{\displaystyle \underset{\alpha =1}{\overset{2}{}}}(p_\alpha +k_\alpha )(h(𝐤)z)^1P_\alpha (0)(h(𝐤)z)^1{\displaystyle \frac{}{}}\}(𝐱,𝐱),`$ $`a_{\mathrm{},1}^{orbit,2}(\omega )`$ $`={\displaystyle \frac{i}{4\pi m^2|\mathrm{\Omega }|}}{\displaystyle _\mathrm{\Omega }}d𝐤{\displaystyle _\mathrm{\Omega }}d𝐱\{\mathrm{tr}{\displaystyle _{\mathrm{\Gamma }_\omega }}dzf_{FD}(z)`$ (7.2) $`\{(p_1+k_1)(h(𝐤)z)^1(p_1+k_1)(h(𝐤)z)^1`$ $`(p_2+k_2)(h(𝐤)z)^1(p_2+k_2)(h(𝐤)z\omega )^1`$ $`+(p_1+k_1)(h(𝐤)z)^1(p_2+k_2)(h(𝐤)z)^1`$ $`(p_1+k_1)(h(𝐤)z)^1(p_2+k_2)(h(𝐤)z\omega )^1`$ $`(p_1+k_1)(h(𝐤)z+\omega )^1(p_1+k_1)(h(𝐤)z+\omega )^1`$ $`(p_2+k_2)(h(𝐤)z+\omega )^1(p_2+k_2)(h(𝐤)z)^1`$ $`+(p_1+k_1)(h(𝐤)z+\omega )^1(p_2+k_2)(h(𝐤)z+\omega )^1`$ $`(p_1+k_1)(h(𝐤)z+\omega )^1(p_2+k_2)(h(𝐤)z)^1`$ $`(p_1+k_1)(h(𝐤)z)^1(p_2+k_2)(h(𝐤)z\omega )^1`$ $`(p_1+k_1)(h(𝐤)z\omega )^1(p_2+k_2)(h(𝐤)z\omega )^1`$ $`+(p_1+k_1)(h(𝐤)z)^1(p_2+k_2)(h(𝐤)z\omega )^1`$ $`(p_2+k_2)(h(𝐤)z\omega )^1(p_1+k_1)(h(𝐤)z\omega )^1`$ $`(p_1+k_1)(h(𝐤)z+\omega )^1(p_2+k_2)(h(𝐤)z)^1`$ $`(p_1+k_1)(h(𝐤)z)^1(p_2+k_2)(h(𝐤)z)^1`$ $`+(p_1+k_1)(h(𝐤)z+\omega )^1(p_2+k_2)(h(𝐤)z)^1`$ $`(p_2+k_2)(h(𝐤)z)^1(p_1+k_1)(h(𝐤)z)^1\}{\displaystyle \frac{}{}}\}(𝐱,𝐱),`$ and $`a_{\mathrm{},1}^{spin}(\omega )={\displaystyle \frac{gc\mu _b}{2e\pi |\mathrm{\Omega }|}}{\displaystyle _\mathrm{\Omega }}d𝐤{\displaystyle _\mathrm{\Omega }}d𝐱\{\mathrm{tr}{\displaystyle _{\mathrm{\Gamma }_\omega }}dzf_{FD}(z)`$ (7.3) $`\{\left[(p_1+k_1)(h(𝐤)z)^1\sigma _3(h(𝐤)z)^1(p_2+k_2)(h(𝐤)z\omega )^1\right]`$ $`+\left[(p_1+k_1)(h(𝐤)z)^1(p_2+k_2)(h(𝐤)z\omega )^1\sigma _3(h(𝐤)z\omega )^1\right]`$ $`+\left[(p_1+k_1)(h(𝐤)z+\omega )^1\sigma _3(h(𝐤)z+\omega )^1(p_2+k_2)(h(𝐤)z)^1\right]`$ $`+\left[(p_1+k_1)(h(𝐤)z+\omega )^1(p_2+k_2)(h(𝐤)z)^1\sigma _3(h(𝐤)z)^1\right]\}{\displaystyle \frac{}{}}\}(𝐱,𝐱).`$ Finally, for the convenience of the reader only interested in applying the theory to the case when one assumes that the Bloch bands and functions are known (as for example from Kohn-Luttinger type models), we write (7.1)-(7.3) in terms of Bloch functions and energies. The important thing here is that no derivatives with respect to the quasi-momentum appear. With the usual notation (here $`,`$ denotes the scalar product over the spin variables): $`\widehat{\pi }_{ij}(\alpha ,𝐤)={\displaystyle _\mathrm{\Omega }}u_i(𝐱,𝐤),(p_\alpha +𝐤_\alpha )u_j(𝐱,𝐤)𝑑𝐱,`$ (7.4) and after some rearrangements, the terms coming from the orbital magnetism are: $`a_{\mathrm{},1}^{orbit,1}(\omega )`$ $`={\displaystyle \frac{1}{2m\omega (2\pi )^3}}{\displaystyle \underset{\alpha =1}{\overset{2}{}}}{\displaystyle _\mathrm{\Omega }^{}}d𝐤\{{\displaystyle \underset{j1}{}}|\widehat{\pi }_{jj}(\alpha ,𝐤)|^2f_{FD}^{}(\lambda _j(𝐤))`$ (7.5) $`\omega ^2{\displaystyle \underset{jl}{}}|\widehat{\pi }_{lj}(\alpha ,𝐤)|^2{\displaystyle \frac{f_{FD}(\lambda _j(𝐤))f_{FD}(\lambda _l(𝐤))}{[(\lambda _j(𝐤)\lambda _l(𝐤))^2\omega ^2](\lambda _j(𝐤)\lambda _l(𝐤))}}\},`$ $`a_{\mathrm{},1}^{orbit,2}(\omega )`$ $`={\displaystyle \frac{1}{2m^2(2\pi )^3}}{\displaystyle _\mathrm{\Omega }^{}}𝑑𝐤{\displaystyle \underset{n_1,n_2,n_3,n_41}{}}{\displaystyle \frac{1}{2\pi i}}{\displaystyle _{\mathrm{\Gamma }_\omega }}𝑑zf_{FD}(z)`$ (7.6) $`\{{\displaystyle \frac{\widehat{\pi }_{n_4n_1}(1,𝐤)\widehat{\pi }_{n_1n_2}(1,𝐤)\widehat{\pi }_{n_2n_3}(2,𝐤)\widehat{\pi }_{n_3n_4}(2,𝐤)}{(z\lambda _{n_1}(𝐤))(z\lambda _{n_2}(𝐤))(z\lambda _{n_3}(𝐤))(z+\omega \lambda _{n_4}(𝐤))}}`$ $`{\displaystyle \frac{\widehat{\pi }_{n_4n_1}(1,𝐤)\widehat{\pi }_{n_1n_2}(2,𝐤)\widehat{\pi }_{n_2n_3}(1,𝐤)\widehat{\pi }_{n_3n_4}(2,𝐤)}{(z\lambda _{n_1}(𝐤))(z\lambda _{n_2}(𝐤))(z\lambda _{n_3}(𝐤))(z+\omega \lambda _{n_4}(𝐤))}}`$ $`+{\displaystyle \frac{\widehat{\pi }_{n_4n_1}(1,𝐤)\widehat{\pi }_{n_1n_2}(1,𝐤)\widehat{\pi }_{n_2n_3}(2,𝐤)\widehat{\pi }_{n_3n_4}(2,𝐤)}{(z\omega \lambda _{n_1}(𝐤))(z\omega \lambda _{n_2}(𝐤))(z\omega \lambda _{n_3}(𝐤))(z\lambda _{n_4}(𝐤))}}`$ $`{\displaystyle \frac{\widehat{\pi }_{n_4n_1}(1,𝐤)\widehat{\pi }_{n_1n_2}(2,𝐤)\widehat{\pi }_{n_2n_3}(1,𝐤)\widehat{\pi }_{n_3n_4}(2,𝐤)}{(z\omega \lambda _{n_1}(𝐤))(z\omega \lambda _{n_2}(𝐤))(z\omega \lambda _{n_3}(𝐤))(z\lambda _{n_4}(𝐤))}}`$ $`+{\displaystyle \frac{\widehat{\pi }_{n_4n_1}(1,𝐤)\widehat{\pi }_{n_1n_2}(2,𝐤)\widehat{\pi }_{n_2n_3}(1,𝐤)\widehat{\pi }_{n_3n_4}(2,𝐤)}{(z\lambda _{n_1}(𝐤))(z+\omega \lambda _{n_2}(𝐤))(z+\omega \lambda _{n_3}(𝐤))(z+\omega \lambda _{n_4}(𝐤))}}`$ $`{\displaystyle \frac{\widehat{\pi }_{n_4n_1}(1,𝐤)\widehat{\pi }_{n_1n_2}(2,𝐤)\widehat{\pi }_{n_2n_3}(2,𝐤)\widehat{\pi }_{n_3n_4}(1,𝐤)}{(z\lambda _{n_1}(𝐤))(z+\omega \lambda _{n_2}(𝐤))(z+\omega \lambda _{n_3}(𝐤))(z+\omega \lambda _{n_4}(𝐤))}}`$ $`+{\displaystyle \frac{\widehat{\pi }_{n_4n_1}(1,𝐤)\widehat{\pi }_{n_1n_2}(2,𝐤)\widehat{\pi }_{n_2n_3}(1,𝐤)\widehat{\pi }_{n_3n_4}(2,𝐤)}{(z\omega \lambda _{n_1}(𝐤))(z\lambda _{n_2}(𝐤))(z\lambda _{n_3}(𝐤))(z\lambda _{n_4}(𝐤))}}`$ $`{\displaystyle \frac{\widehat{\pi }_{n_4n_1}(1,𝐤)\widehat{\pi }_{n_1n_2}(2,𝐤)\widehat{\pi }_{n_2n_3}(2,𝐤)\widehat{\pi }_{n_3n_4}(1,𝐤)}{(z\omega \lambda _{n_1}(𝐤))(z\lambda _{n_2}(𝐤))(z\lambda _{n_3}(𝐤))(z\lambda _{n_4}(𝐤))}}\}.`$ As for the spin contribution, with the notation $$\widehat{s}_{ij}(𝐤):=_\mathrm{\Omega }u_i(𝐱,𝐤),\sigma _3u_j(𝐱,𝐤)𝑑𝐱,$$ (7.7) one has: $`a_{\mathrm{},1}^{spin}(\omega )`$ $`={\displaystyle \frac{gc\mu _b}{(2\pi )^4e}}{\displaystyle _\mathrm{\Omega }^{}}𝑑𝐤{\displaystyle \underset{n_1,n_2,n_31}{}}{\displaystyle \frac{1}{2\pi i}}{\displaystyle _{\mathrm{\Gamma }_\omega }}𝑑zf_{FD}(z)`$ (7.8) $`\{{\displaystyle \frac{\widehat{\pi }_{n_1n_2}(1,𝐤)\widehat{s}_{n_2n_3}(𝐤)\widehat{\pi }_{n_3n_1}(2,𝐤)}{(\lambda _{n_2}(𝐤)z)(\lambda _{n_3}(𝐤)z)(\lambda _{n_1}(𝐤)z\omega )}}`$ $`+{\displaystyle \frac{\widehat{\pi }_{n_1n_2}(1,𝐤)\widehat{\pi }_{n_2n_3}(2,𝐤)\widehat{s}_{n_3n_1}(𝐤)}{(\lambda _{n_2}(𝐤)z)(\lambda _{n_3}(𝐤)z\omega )(\lambda _{n_1}(𝐤)z\omega )}}`$ $`+{\displaystyle \frac{\widehat{\pi }_{n_1n_2}(1,𝐤)\widehat{s}_{n_2n_3}(𝐤)\widehat{\pi }_{n_3n_1}(2,𝐤)}{(\lambda _{n_2}(𝐤)z+\omega )(\lambda _{n_3}(𝐤)z+\omega )(\lambda _{n_1}(𝐤)z)}}`$ $`+{\displaystyle \frac{\widehat{\pi }_{n_1n_2}(1,𝐤)\widehat{\pi }_{n_2n_3}(2,𝐤)\widehat{s}_{n_3n_1}(𝐤)}{(\lambda _{n_2}(𝐤)z+\omega )(\lambda _{n_3}(𝐤)z)(\lambda _{n_1}(𝐤)z)}}\}.`$ ## 8 Conclusions We presented in the present paper a method which shed new light on the quantum dynamics/optical response in bulk media in the presence of a constant magnetic field. We applied the gauge invariant magnetic perturbation theory and gave a clear and very general way of dealing with long range magnetic perturbations. The formal connection with the integer Quantum Hall effect was established in (4.2). Equations (6.9)-(6.12) and (7.4)-(7.8) contain our main result concerning the Verdet constant and the Faraday effect: it gives the linear term in $`B`$ of the transverse conductivity in terms of the zero magnetic field Green function. They open the way of using the recently developed Green function techniques for the calculation of optical and magneto-optical properties of solids, to the case when an external magnetic field is present. Our method can be applied to ordered, as well as to random systems (with the appropriate average over configurations). Of course, in the last case one has to assume ergodicity properties in order to insure convergence of results in the thermodynamic limit. Layers or other geometries can also be considered. There are many subtle and difficult mathematical questions left aside in this paper, as those related to the thermodynamic limit, the convergence of infinite series over Bloch bands, and the low frequency limit when the Fermi energy lies in the spectrum. Another open problem is to consider self-interacting electrons and to investigate the exciton influence on the Faraday effect. These questions will be addressed elsewhere. Our results are not only theoretical. In a future publication we will use the residue theorem in equations (7.4)-(7.8) to calculate the Verdet constant for various finite band models, and compare our results with the existing experimental data. Moreover, our results will be shown to imply those of Roth and Nedoluha . ## 9 Acknowledgments This paper has been written at the Department of Mathematical Sciences, Aalborg University. G.N. kindly thanks the Department for invitation and financial support. H.C. acknowledges support from the Statens Naturvidenskabelige Forskningsråd grant Topics in Rigorous Mathematical Physics, and partial support through the European Union’s IHP network Analysis $`\&`$ Quantum HPRN-CT-2002-00277.
warning/0506/math0506430.html
ar5iv
text
# Invariant ideals and Matsushima’s criterion ## 1. Algebraic homogeneous spaces Let $`G`$ be an affine algebraic group over an algebraically closed field $`𝕂`$. A $`G`$-module $`V`$ is said to be rational if any vector in $`V`$ is contained in a finite-dimensional rational $`G`$-submodule. Below all modules are supposed to be rational. By $`V^G`$ denote the subspace of $`G`$-fixed vectors in $`V`$. The group $`G\times G`$ acts on $`G`$ by translations, $`(g_1,g_2)g:=g_1gg_2^1`$. This action induces the action on the algebra of regular functions on $`G`$: $$(G\times G):𝕂[G],((g_1,g_2)f)(g):=f(g_1^1gg_2).$$ For any closed subgroup $`H`$ of $`G`$, $`H_l`$, $`H_r`$ denote the groups of all left and right translations of $`𝕂[G]`$ by elements of $`H`$. Under these actions, the algebra $`𝕂[G]`$ becomes a rational $`H_l`$\- (and $`H_r`$-) module. By Chevalley’s Theorem, the set $`G/H`$ of left $`H`$-cosets in $`G`$ admits a structure of a quasi-projective algebraic variety such that the projection $`p:GG/H`$ is a surjective $`G`$-equivariant morphism. Moreover, a structure of an algebraic variety on $`G/H`$ satisfying these conditions is unique. It is easy to check that the morphism $`p`$ is open and the algebra of regular functions on $`G/H`$ may be identified with the subalgebra $`𝕂[G]^{H_r}`$ in $`𝕂[G]`$. We refer to \[6, Ch. IV\] for details. ## 2. Matsushima’s criterion Let $`G`$ be a reductive algebraic group and $`H`$ a closed subgroup of $`G`$. It is known that the homogeneous space $`G/H`$ is affine if and only if $`H`$ if reductive. The first proof was given over the field of complex numbers and used some results from algebraic topology, see and \[9, Th. 4\]. An algebraic proof in characteristic zero was obtained in . A characteristic-free proof that uses the Mumford conjecture proved by W.J. Haboush is given in . Another proof based on the Morozov-Jacobson Theorem may be found in . Below we give an elementary proof of Matsushima’s criterion in terms of representation theory. The ground field $`𝕂`$ is assumed to be algebraically closed and of characteristic zero. ###### Theorem 2.1. Let $`G`$ be a reductive algebraic group and $`H`$ its closed subgroup. Then the homogeneous space $`G/H`$ is affine if and only if $`H`$ is reductive. ###### Proof. We begin with the “easy half”. ###### Proposition 2.2. Let $`G`$ be an affine algebraic group and $`H`$ its reductive subgroup. Then $`G/H`$ is affine. ###### Proof. If a reductive group $`H`$ acts on an affine variety $`X`$, then the algebra of invariants $`𝕂[X]^H`$ is finitely generated, the quotient morphism $`\pi :X\mathrm{Spec}𝕂[X]^H`$ is surjective and any fiber of $`\pi `$ contains a unique closed $`H`$-orbit \[10, Sec .4.4\]. In the case $`X=G`$ this shows that $`G/H`$ is isomorphic to $`\mathrm{Spec}𝕂[G]^{H_r}`$. Now assume that $`G`$ is reductive and consider a decomposition $$𝕂[G]=𝕂𝕂[G]_G,$$ where the first component corresponds to constant functions on $`G`$, and the second one is the sum of all simple non-trivial $`G_l`$\- (or $`G_r`$-) submodules in $`𝕂[G]`$. Let $`\mathrm{pr}:𝕂[G]𝕂`$ be the projection on the first component. Clearly, $`\mathrm{pr}`$ is a $`(G_l\times G_r)`$-invariant linear map. Let $`H`$ be a closed subgroup of $`G`$. Consider $$I(G,H)=\{f𝕂[G]^{H_r}|\mathrm{pr}(fl)=0l𝕂[G]^{H_r}\}.$$ This is a $`G_l`$-invariant ideal in $`𝕂[G]^{H_r}`$ with $`1I(G,H)`$. Assume that $`G/H`$ is affine. Then $`G/H\mathrm{Spec}𝕂[G]^{H_r}`$ and $`𝕂[G]^{H_r}`$ does not contain proper $`G_l`$-invariant ideals. Thus $`I(G,H)=0`$. Our aim is to deduce from this that any $`H`$-module is completely reducible. ###### Lemma 2.3. If $`W`$ is an $`H_r`$-submodule in $`𝕂[G]`$ and $`fW`$ is a non-zero $`H_r`$-fixed vector, then $`W=fW^{}`$, where $`W^{}`$ is an $`H_r`$-submodule. ###### Proof. Since $`I(G,H)=0`$, there exists $`l𝕂[G]^{H_r}`$ such that $`\mathrm{pr}(fl)0`$. The submodule $`W^{}`$ is defined as $`W^{}=\{wW|\mathrm{pr}(wl)=0\}`$. ###### Lemma 2.4. If $`f𝕂[G]`$ is an $`H_r`$-semi-invariant of the weight $`\xi `$, then there exists an $`H_r`$-semi-invariant in $`𝕂[G]`$ of the weight $`\xi `$. ###### Proof. Let $`Z`$ be the zero set of $`f`$ in $`G`$. Since $`Z`$ is $`H_r`$-invariant, one has $`Z=p^1(p(Z))`$. This implies that $`p(Z)`$ is a proper closed subset of $`G/H`$. There exists a non-zero $`\alpha 𝕂[G/H]`$ with $`\alpha |_{p(Z)}=0`$. Then $`p^{}\alpha 𝕂[G]^{H_r}`$ and $`p^{}\alpha |_Z=0`$. By Hilbert’s Nullstellensatz, there are $`n`$, $`s𝕂[G]`$ such that $`(p^{}\alpha )^n=fs`$. This shows that $`s`$ is an $`H_r`$-semi-invariant of the weight $`\xi `$. ###### Lemma 2.5. (1) Any cyclic $`G`$-module $`V`$ may be embedded (as a $`G_r`$-submodule) into $`𝕂[G]`$. (2) Any $`n`$-dimensional $`H`$-module $`W`$ may be embedded (as an $`H_r`$-submodule) into $`(𝕂[H])^n`$. (3) Any finite-dimensional $`H`$-module may be embedded (as an $`H`$-submodule) into a finite-dimensional $`G`$-module. ###### Proof. (1) Suppose that $`V=Gv`$. The map $`\varphi :GV`$, $`\varphi (g)=g^1v`$, induces the embedding of the dual module $`\varphi ^{}:V^{}𝕂[G]`$. Consider the $`G_r`$-submodule $`U=\{f𝕂[G]|\mathrm{pr}(fl)=0l\varphi ^{}(V^{})\}`$. By the complete reducibility, $`𝕂[G]=UU^{}`$ for some $`G_r`$-submodule $`U^{}`$. Obviously, $`I(G,G)=0`$ and $`U^{}`$ is $`G_r`$-isomorphic to $`V`$. (2) Let $`\lambda _1,\mathrm{},\lambda _n`$ be a basis of $`W^{}`$. The embedding may be given as $$w(f_1^w,\mathrm{},f_n^w),f_i^w(h):=\lambda _i(hw).$$ (3) Note that the restriction homomorphism $`𝕂[G]𝕂[H]`$ is surjective. By (2), any finite-dimensional $`H`$-module $`W`$ has the form $`W_1/W_2`$, where $`W_1`$ is a finite-dimensional $`H`$-submodule in a $`G`$-module $`V`$ and $`W_2`$ is an $`H`$-submodule of $`W_1`$. Consider $`W_1(^mW_2)`$ as an $`H`$-submodule in $`^{m+1}W_1`$, where $`m=dimW_2`$. Note that $`W(W_1(^mW_2))(^mW_2)^{}`$. By (1), the cyclic $`G`$-submodule of $`^mV`$ generated by $`^mW_2`$ may be embedded into $`𝕂[G]`$. By Lemma 2.4, $`(^mW_2)^{}`$ also may be embedded into a $`G`$-module. ###### Lemma 2.6. For any $`H`$-module $`W`$ and any non-zero $`wW^H`$ there is an $`H`$-submodule $`W^{}`$ such that $`W=wW^{}`$. ###### Proof. Embed $`W`$ into a $`G`$-module $`V`$. Let $`V_1=Gw`$. Then $`V=V_1V_2`$ for some $`G`$-submodule $`V_2`$. Embed $`V_1`$ into $`𝕂[G]`$ as a $`G_r`$-submodule. By Lemma 2.3, $`V_1=wW_1`$ for some $`H`$-submodule $`W_1`$. Finally, $`W^{}=W(W_1V_2)`$. ###### Lemma 2.7. Any $`H`$-module is completely reducible. ###### Proof. Assume that $`W_1`$ is a simple submodule in an $`H`$-module $`W`$. Consider two submodules in the $`H`$-module $`\mathrm{End}(W,W_1)`$: $$L_2=\{p\mathrm{End}(W,W_1)|p|_{W_1}=0\}L_1=\{p\mathrm{End}(W,W_1)|p|_{W_1}\mathrm{is}\mathrm{scalar}\}.$$ Clearly, $`L_2`$ is a hyperplane in $`L_1`$. Consider an $`H`$-eigenvector $`l(L_1)^{}`$ corresponding to $`L_2`$. Taking the tensor product with a one-dimensional $`H`$-module, one may assume that $`l`$ is $`H`$-fixed. By Lemma 2.6, $`(L_1)^{}=lM`$, implying $`L_1=L_2P`$, where $`M`$ and $`P`$ are $`H`$-submodules. Then $`\mathrm{Ker}P`$ is a complementary submodule to $`W_1`$. Theorem 2.1 is proved. ###### Remark 2.8. In , for any action $`G:X`$ of a reductive group $`G`$ on an affine variety $`X`$ with the decomposition $`𝕂[X]=𝕂[X]^G𝕂[X]_G`$ and the projection $`\mathrm{pr}:𝕂[X]𝕂[X]^G`$, the $`𝕂[X]^G`$-bilinear scalar product $`(f,g)=\mathrm{pr}(fg)`$ on $`𝕂[X]`$ was introduced and the kernel of this product was considered. Our ideal $`I(G,H)`$ is such kernel in the case $`X=\mathrm{Spec}𝕂[G]^{H_r}`$ provided $`𝕂[G]^{H_r}`$ is finitely generated. ###### Remark 2.9. For convenience of the reader we include all details in the proof of Theorem 2.1. Lemma 2.4 and Lemma 2.5 are taken from . They show that for a quasi-affine $`G/H`$ any $`H`$-module may be realized as an $`H`$-submodule of a $`G`$-module. The converse is also true . Proposition 2.2 is a standart fact. The proof of Lemma 2.7 is a part of the proof of the Weyl Theorem on complete reducibility , see also \[12, Prop. 2.2.4\]. ## 3. Some additional remarks The following lemma may be found in . ###### Lemma 3.1. Let $`G`$ be an affine algebraic group and $`H`$ its reductive subgroup. Then $`𝕂[G]^{H_r}`$ does not contain proper $`G_l`$-invariant ideals. ###### Proof. Consider a decomposition $$𝕂[G]=𝕂[G]^{H_r}𝕂[G]_{H_r},$$ where $`𝕂[G]_{H_r}`$ is the sum of all non-trivial simple $`H_r`$-submodules in $`𝕂[G]`$. Clearly, $`𝕂[G]^{H_r}𝕂[G]_{H_r}𝕂[G]_{H_r}`$. Hence any proper $`G_l`$-invariant ideal in $`𝕂[G]^{H_r}`$ generates a proper $`G_l`$-invariant ideal in $`𝕂[G]`$, a contradiction. By Hilbert’s Theorem on invariants, the algebra $`𝕂[G]^{H_r}`$ is finitely generated. It is easy to see that functions from $`𝕂[G]^{H_r}`$ separate (closed) right $`H`$-cosets in $`G`$. These observations and Lemma 3.1 give another proof of Proposition 2.2. Moreover, it is proved in \[2, Prop. 1\] that for a quasi-affine $`G/H`$ the algebra $`𝕂[G]^{H_r}`$ does not contain proper $`G_l`$-invariant ideals if and only if $`G/H`$ is affine. Now assume that $`G`$ is reductive. ###### Proposition 3.2. \[13, Prop. 1\] The ideal $`I(G,H)`$ is the biggest $`G_l`$-invariant ideal in $`𝕂[G]^{H_r}`$ different from $`𝕂[G]^{H_r}`$. ###### Proof. Any proper $`G_l`$-invariant ideal $`I`$ of $`𝕂[G]^{H_r}`$ is contained in $`𝕂[G]^{H_r}𝕂[G]_G`$. Thus $`\mathrm{pr}(il)=0`$ for any $`l𝕂[G]^{H_r}`$, $`iI`$. This implies $`II(G,H)`$. ###### Remark 3.3. For non-reductive $`G`$ the biggest invariant ideal in $`𝕂[G]^{H_r}`$ may not exist. For example, one may take $$G=\left\{\left(\begin{array}{ccc}1& & \\ 0& & \\ 0& & \end{array}\right)\right\},H=\left\{\left(\begin{array}{ccc}1& 0& \\ 0& 1& \\ 0& 0& \end{array}\right)\right\}.$$ Here $`G/H𝕂^3\{x_2=x_3=0\}`$, $`𝕂[G]^{H_r}𝕂[x_1,x_2,x_3]`$, and the maximal ideals $`(x_1a,x_2,x_3)`$ are $`G_l`$-invariant for any $`a𝕂`$. ## 4. The boundary ideal In this section we assume that $`H`$ is an observable subgroup of $`G`$, i.e., $`G/H`$ is quasi-affine. If the algebra $`𝕂[G]^{H_r}`$ is finitely generated, then the affine $`G`$-variety $`X=\mathrm{Spec}𝕂[G]^{H_r}`$ has an open $`G`$-orbit isomorphic to $`G/H`$ and may be considered as the canonical embedding $`G/HX`$. Moreover, this embedding is uniquely characterized by two properties: $`X`$ is normal and $`\mathrm{codim}_X(XG/H)2`$, see . There are two remarkable $`G_l`$-invariant ideals in $`𝕂[G]^{H_r}`$, namely $$I^b(G,H)=I(X(G/H))=\{f𝕂[G]^{H_r}|f|_{X(G/H)}=0\},$$ and, if $`G`$ is reductive, the ideal $`I^m(G,H)`$ of the unique closed $`G`$-orbit in $`X`$. If $`G/H`$ is affine, then $`I^b(G,H)=𝕂[G]^{H_r}`$, $`I^m(G,H)=0`$. In other cases $`I^b(G,H)`$ is the smallest proper radical $`G_l`$-invariant ideal, and $`I^m(G.H)`$ is the biggest proper $`G_l`$-invariant ideal of $`𝕂[G]^{H_r}`$. By Proposition 3.2, $`I^m(G,H)=I(G,H)`$. Moreover, $`𝕂[G]^{H_r}/I^m(G,H)𝕂[G]^{S_r}`$, where $`S`$ is a minimal reductive subgroup of $`G`$ containing $`H`$. (Such a subgroup may be not unique, but all of them are $`G`$-conjugate, see \[1, Sec. 7\].) It follows from the Slice Theorem and \[1, Prop. 4\] that $`I^b(G,H)=I^m(G,H)`$ if and only if $`H`$ is a quasi-parabolic subgroup of a reductive subgroup of $`G`$. Now assume that $`𝕂[G]^{H_r}`$ is not finitely generated. If $`G`$ is reductive, then $`I(G,H)`$ may be consider as an analog of $`I^m(G,H)`$ in this situation (Proposition 3.2). We claim that $`I^b(G,H)`$ also has an analog, even for non-reductive $`G`$. ###### Proposition 4.1. Let $`\widehat{X}`$ be a quasi-affine variety, $`\widehat{X}X`$ be an (open) embedding into an affine variety $`X`$, $`I(X\widehat{X})𝕂[X]`$, and $`=(\widehat{X})`$ be the radical of the ideal of $`𝕂[\widehat{X}]`$ generated by $`I(X\widehat{X})`$. Then (1) the ideal $`𝕂[\widehat{X}]`$ does not depend on $`X`$; (2) $`I(X\widehat{X})`$ is the smallest radical ideal of $`𝕂[X]`$ generating an ideal in $`𝕂[\widehat{X}]`$ with the radical $``$. ###### Proof. (1) Consider two affine embeddings: $`\varphi _i:\widehat{X}X_i`$, $`i=1,2`$. Let $`X_{12}`$ be the closure of $`(\varphi _1\times \varphi _2)(\widehat{X})`$ in $`X_1\times X_2`$ with the projections $`r_i:X_{12}X_i`$. Let us identify the images of $`\widehat{X}`$ in $`X_1`$, $`X_2`$, and $`X_{12}`$. We claim that $`r_i(X_{12}\widehat{X})X_i\widehat{X}`$. Indeed, the diagonal image of $`\widehat{X}`$ is closed in $`\widehat{X}\times X_j`$, $`ji`$, as the graph of a morphism. It follows from what was proved above that the ideal of $`𝕂[X_{12}]`$ generated by $`r_i^{}(I(X_i\widehat{X}))`$ has the radical $`I(X_{12}\widehat{X})`$. This shows that the radical of the ideal generated by $`I(X_i\widehat{X})`$ in $`𝕂[\widehat{X}]`$ does not depend on $`i`$. (2) Assume that there is a radical ideal $`I_1𝕂[X]`$ not containing $`I=I(X\widehat{X})`$ and generating an ideal in $`𝕂[\widehat{X}]`$ with the radical $``$. There is $`x_0\widehat{X}`$ such that $`h(x_0)=0`$ for any $`hI_1`$. Take $`fI`$ such that $`f(x_0)0`$. One has $`f^k=\alpha _1h_1+\mathrm{}+\alpha _kh_k`$ for some $`\alpha _i𝕂[\widehat{X}]`$, $`h_iI_1`$, $`k`$, and this implies $`f(x_0)=0`$, a contradiction. So $`(G/H)`$ is a radical $`G_l`$-invariant ideal of $`𝕂[G]^{H_r}`$, and $`(G/H)=I^b(G,H)`$ provided $`𝕂[G]^{H_r}`$ is finitely generated. ###### Proposition 4.2. $`(G/H)`$ is the smallest non-zero radical $`G_l`$-invariant ideal of $`𝕂[G]^{H_r}`$. ###### Proof. Let $`f𝕂[G]^{H_r}`$ and $`I(f)`$ be the ideal of $`𝕂[G]^{H_r}`$ generated by the orbit $`G_lf`$. It is sufficient to prove that $`(G/H)\mathrm{rad}I(f)`$. Take any $`G`$-equivariant affine embedding $`G/HX`$ with $`f𝕂[X]`$. For the ideal $`I^{}(f)`$ generated by $`G_lf`$ in $`𝕂[X]`$ one has $`I(X(G/H))\mathrm{rad}I^{}(f)`$, hence $`(G/H)\mathrm{rad}I(f)`$. ###### Corollary 4.3. Let $`G`$ be an affine algebraic group and $`H`$ its observable subgroup. Then $`G/H`$ is affine if and only if $`(G/H)=𝕂[G]^{H_r}`$. It should be interesting to give a description of the ideal $`(G/H)`$ similar to the definition of $`I(G,H)`$, and to find a geometric meaning of the $`G_l`$-algebras $`𝕂[G]^{H_r}/I(G,H)`$ and $`𝕂[G]^{H_r}/(G/H)`$ for non-finitely generated $`𝕂[G]^{H_r}`$. Acknowledgements. The author is grateful to J. Hausen for useful discussions. In particular, Proposition 4.1 appears during such a discussion. Thanks are also due to D.A. Timashev for valuable remarks. This paper was written during the staying at Eberhard Karls Universität Tübingen (Germany). The author wishes to thank this institution and especially Jürgen Hausen for invitation and hospitality.
warning/0506/math-ph0506003.html
ar5iv
text
# EXTENDED HAMILTONIAN SYSTEMS IN MULTISYMPLECTIC FIELD THEORIES ## 1 Introduction The Hamiltonian formalism of dynamical systems, and the study of the properties of Hamiltonian dynamical systems in general, is a fruitful subject in both applied mathematics and theoretical physics. From a generic point of view, the characteristics of these kinds of systems make them specially suitable for analyzing many of their properties; for instance: symmetries and related topics such as the existence of conservation laws and reduction, the integrability (including numerical methods), and the possible quantization of the system, which is based on the use of the Poisson bracket structure of this formalism. Moreover, it is also important to point out the existence of dynamical Hamiltonian systems which have no Lagrangian counterpart (see an example in ). From the geometrical viewpoint, many of the characteristics of the autonomous Hamiltonian systems arise from the existence of a “natural” geometric structure with which the phase space of the system is endowed: the symplectic form (a closed, nondegenerated two-form), which allows the construction of Poisson brackets. In this model, the dynamic information is carried out by the Hamiltonian function, which is not coupled to the geometry. This is not the case for non-autonomous Hamiltonian systems, which have different geometric descriptions. One of the most frequently used formulations for these systems is in the framework of contact geometry, which takes place in the restricted phase space $`\mathrm{T}^{}Q\times `$, where $`Q`$ is the configuration manifold (see and references therein). Here, the physical information is given by the Hamiltonian function, which allows us to construct the contact form in $`\mathrm{T}^{}Q\times `$. However, a more appropriate description is the symplectic or extended formulation of non-autonomous mechanics , , , , , which is developed in the extended phase space $`\mathrm{T}^{}(Q\times )`$. Now, the natural symplectic structure of $`\mathrm{T}^{}(Q\times )`$ and the physical information, given by the extended Hamiltonian function, are decoupled and this provides us with a Hamiltonian description similar to the autonomous case. When first-order field theories are considered, the usual way to work is with the Lagrangian formalism , , , , , , , , , because their Hamiltonian description presents different kinds of problems. First, several Hamiltonian models can be stated, and the equivalence among them is not always clear (see, for instance, , , , , , , , ). Furthermore, there are equivalent Lagrangian models with non-equivalent Hamiltonian descriptions , , . Among the different geometrical descriptions to be considered for describing field theories, we focus our attention on the multisymplectic models , , , , ; where the geometric background is in the realm of multisymplectic manifolds, which are manifolds endowed with a closed and $`1`$-nondegenerate $`k`$-form, with $`k2`$. In these models, this form plays a similar role to the symplectic form in mechanics. The main aim of this paper is to generalize the Hamiltonian symplectic formulation of non-autonomous mechanics to first-order multisymplectic field theories. The motivation and basic features of this formulation are the following: As is well known, there is no a canonical model for Hamiltonian first-order field theory. Hence the first problem to be considered is the choice of a suitable multimomentum bundle to develop the formalism. The most frequently used choice is to take the so-called restricted multimomentum bundle, denoted by $`J^1\pi ^{}`$; that is analogous to $`\mathrm{T}^{}Q\times `$ in the mechanical case. The Hamiltonian formalism in $`J^1\pi ^{}`$ has been extensively studied , , , . Nevertheless, this bundle does not have a canonical multisymplectic form and the physical information, given by a Hamiltonian section, is used to obtain the geometric structure. This is a problem when other aspects of Hamiltonian field theories are considered, such as: the definition of Poisson brackets, the notion of integrable system, the problem of reduction by symmetries, and the quantization procedure. An attempt to overcome these difficulties is to work in a greater dimensional manifold, the so-called extended multimomentum bundle, denoted by $`\pi `$, which is the analogous to the extended phase space $`\mathrm{T}^{}(Q\times )`$ of a non-autonomous mechanical system. $`\pi `$ has a canonical multisymplectic form, since it is a vector subbundle of a multicotangent bundle. In this manifold $`\pi `$, the physical information is given by a closed one form, the Hamiltonian form. Then Hamiltonian systems can be introduced as in autonomous mechanics, by using certain kinds of Hamiltonian multivector fields. The resultant extended Hamiltonian formalism is the generalization to field theories of the extended formalism for non-autonomous mechanical systems , and, to our knowledge, it was introduced for the first time in field theories in . The goal of our work is to carry out a deeper geometric study of these kinds of systems. The main results are the following: first, to every Hamiltonian system in the extended multimomentum bundle, we can associate in a natural way a class of equivalent Hamiltonian systems in the restricted multimomentum bundle (Theorem 5), and conversely (Proposition 7). The solutions to the field equations in both models are also canonically related. In addition, the field equations for these kinds of systems can be derived from an appropriate variational principle (Theorem 6), which constitutes a first attempt to tackle variational principles for field theories with non-holonomic constraints (see for a geometrical setting of these theories). Furthermore, the integral submanifolds of the Hamiltonian $`1`$-form can be embedded into the extended multimomentum phase space similar to the way in which the constant energy surfaces are coisotropically embedded in $`\mathrm{T}^{}(Q\times )`$ in non-autonomous mechanics (Proposition 6). Finally, the case of non regular Hamiltonian systems is considered and, after a carefull definition of what an almost-regular Hamiltonian system is, the above results are adapted to this situation in a natural way. We hope that all these results could be a standpoint from which to study Poisson brackets, the quantization problem and also the reduction by symmetries of field theories in further research works. The paper is organized as follows: In Section 2 we review basic concepts and results, such as multivector fields and connections, multisymplectic manifolds and Hamiltonian multivector fields, and the restricted and extended multimomentum bundles with their geometric structures. Section 3 is devoted to reviewing the definition and characteristics of Hamiltonian systems in the restricted multimomentum bundles; in particular, the definitions of Hamiltonian sections and densities, the variational principle which leads to Hamilton-De Donder-Weyl equations, and the use of multivector fields for writing these equations in a more suitable geometric way. Sections 4 and 5 contain the most relevant material of the work. Thus, Hamiltonian systems in the extended multimomentum bundle are introduced in Section 4; in particular, their geometric properties, their relation with those introduced in Section 3, and the corresponding variational principle are studied. In Section 5 we adapt the above definitions and results in order to consider Hamiltonian systems which are not defined everywhere in the multimomentum bundles, but in certain submanifolds of them: here, these are the so-called almost regular Hamiltonian systems. Finally, as typical examples, in Section 6, we review the standard Hamiltonian formalism associated to a Lagrangian field theory, both in the regular and singular (almost-regular) cases, and the Hamiltonian formalisms of time-dependent dynamical systems in the extended and restricted phase space, which are recovered as a particular case of this theory. All manifolds are real, paracompact, connected and $`C^{\mathrm{}}`$. All maps are $`C^{\mathrm{}}`$. Sum over crossed repeated indices is understood. Throughout this paper $`\pi :EM`$ will be a fiber bundle ($`dimM=m`$, $`dimE=n+m`$), where $`M`$ is an oriented manifold with volume form $`\omega \mathrm{\Omega }^m(M)`$, and $`(x^\nu ,y^A)`$ (with $`\nu =1,\mathrm{},m`$; $`A=1,\mathrm{},n`$) will be natural local systems of coordinates in $`E`$ adapted to the bundle, such that $`\omega =\mathrm{d}x^1\mathrm{}\mathrm{d}x^m\mathrm{d}^mx`$. ## 2 Previous definitions and results ### 2.1 Multivector fields and connections (See for details). Let $``$ be a $`n`$-dimensional differentiable manifold. Sections of $`\mathrm{\Lambda }^m(\mathrm{T})`$ are called $`m`$-multivector fields in $``$ (they are the contravariant skew-symmetric tensors of order $`m`$ in $``$). We will denote by $`\text{X}^m()`$ the set of $`m`$-multivector fields in $``$. If $`𝒴\text{X}^m()`$, for every $`p`$, there exists an open neighbourhood $`U_p`$ and $`Y_1,\mathrm{},Y_r\text{X}(U_p)`$ such that $`𝒴\underset{U_p}{=}{\displaystyle \underset{1i_1<\mathrm{}<i_mr}{}}f^{i_1\mathrm{}i_m}Y_{i_1}\mathrm{}Y_{i_m}`$, with $`f^{i_1\mathrm{}i_m}\mathrm{C}^{\mathrm{}}(U_p)`$ and $`mr\mathrm{dim}`$. Then, $`𝒴\text{X}^m()`$ is said to be locally decomposable if, for every $`p`$, there exists an open neighbourhood $`U_p`$ and $`Y_1,\mathrm{},Y_m\text{X}(U_p)`$ such that $`𝒴\underset{U_p}{=}Y_1\mathrm{}Y_m`$. A non-vanishing $`m`$-multivector field $`𝒴\text{X}^m()`$ and a $`m`$-dimensional distribution $`D\mathrm{T}`$ are locally associated if there exists a connected open set $`U`$ such that $`𝒴|_U`$ is a section of $`\mathrm{\Lambda }^mD|_U`$. If $`𝒴,𝒴^{}\text{X}^m()`$ are non-vanishing multivector fields locally associated with the same distribution $`D`$, on the same connected open set $`U`$, then there exists a non-vanishing function $`f\mathrm{C}^{\mathrm{}}(U)`$ such that $`𝒴^{}\underset{U}{=}f𝒴`$. This fact defines an equivalence relation in the set of non-vanishing $`m`$-multivector fields in $``$, whose equivalence classes will be denoted by $`\{𝒴\}_U`$. Then there is a one-to-one correspondence between the set of $`m`$-dimensional orientable distributions $`D`$ in $`\mathrm{T}`$ and the set of the equivalence classes $`\{𝒴\}_{}`$ of non-vanishing, locally decomposable $`m`$-multivector fields in $``$. If $`𝒴\text{X}^m()`$ is non-vanishing and locally decomposable, and $`U`$ is a connected open set, the distribution associated with the class $`\{𝒴\}_U`$ is denoted by $`𝒟_U(𝒴)`$. If $`U=`$ we write $`𝒟(𝒴)`$. A non-vanishing, locally decomposable multivector field $`𝒴\text{X}^m()`$ is said to be integrable (resp. involutive) if its associated distribution $`𝒟_U(𝒴)`$ is integrable (resp. involutive). Of course, if $`𝒴\text{X}^m()`$ is integrable (resp. involutive), then so is every other in its equivalence class $`\{𝒴\}`$, and all of them have the same integral manifolds. Moreover, Frobenius theorem allows us to say that a non-vanishing and locally decomposable multivector field is integrable if, and only if, it is involutive. Nevertheless, in many applications we have locally decomposable multivector fields $`𝒴\text{X}^m()`$ which are not integrable in $``$, but integrable in a submanifold of $``$. A (local) algorithm for finding this submanifold has been developed . The particular situation in which we are interested is the study of multivector fields in fiber bundles. If $`\pi :M`$ is a fiber bundle, we will be interested in the case where the integral manifolds of integrable multivector fields in $``$ are sections of $`\pi `$. Thus, $`𝒴\text{X}^m()`$ is said to be $`\pi `$-transverse if, at every point $`y`$, $`(𝑖(𝒴)(\pi ^{}\beta ))_y0`$, for every $`\beta \mathrm{\Omega }^m(M)`$ with $`\beta (\pi (y))0`$. Then, if $`𝒴\text{X}^m()`$ is integrable, it is $`\pi `$-transverse if, and only if, its integral manifolds are local sections of $`\pi :M`$. In this case, if $`\varphi :UM`$ is a local section with $`\varphi (x)=y`$ and $`\varphi (U)`$ is the integral manifold of $`𝒴`$ through $`y`$, then $`\mathrm{T}_y(\mathrm{Im}\varphi )=𝒟_y(𝒴)`$. Finally, it is clear that classes of locally decomposable and $`\pi `$-transverse multivector fields $`\{𝒴\}\text{X}^m()`$ are in one-to-one correspondence with orientable Ehresmann connection forms $``$ in $`\pi :M`$. This correspondence is characterized by the fact that the horizontal subbundle associated with $``$ is $`𝒟(𝒴)`$. In this correspondence, classes of integrable locally decomposable and $`\pi `$-transverse $`m`$ multivector fields correspond to flat orientable Ehresmann connections. ### 2.2 Hamiltonian multivector fields in multisymplectic manifolds (See and for details). Let $``$ be a $`n`$-dimensional differentiable manifold and $`\mathrm{\Omega }\mathrm{\Omega }^{m+1}()`$. The couple $`(,\mathrm{\Omega })`$ is said to be a multisymplectic manifold if $`\mathrm{\Omega }`$ is closed and $`1`$-nondegenerate; that is, for every $`p`$, and $`X_p\mathrm{T}_p`$, we have that $`𝑖(X_p)\mathrm{\Omega }_p=0`$ if, and only if, $`X_p=0`$. If $`(,\mathrm{\Omega })`$ is a multisymplectic manifold, $`𝒳\text{X}^k()`$ is said to be a Hamiltonian $`k`$-multivector field if $`𝑖(𝒳)\mathrm{\Omega }`$ is an exact $`(m+1k)`$-form; that is, there exists $`\zeta \mathrm{\Omega }^{mk}()`$ such that $$𝑖(𝒳)\mathrm{\Omega }=\mathrm{d}\zeta $$ (1) $`\zeta `$ is defined modulo closed $`(mk)`$-forms. The class $`\{\zeta \}\mathrm{\Omega }^{mk}()/Z^{mk}()`$ defined by $`\zeta `$ is called the Hamiltonian for $`𝒳`$, and every element in this class $`\widehat{\zeta }\{\zeta \}`$ is said to be a Hamiltonian form for $`𝒳`$. Furthermore, $`𝒳`$ is said to be a locally Hamiltonian $`k`$-multivector field if $`𝑖(𝒳)\mathrm{\Omega }`$ is a closed $`(m+1k)`$-form. In this case, for every point $`x`$, there is an open neighbourhood $`W`$ and $`\zeta \mathrm{\Omega }^{mk}(W)`$ such that $$𝑖(𝒳)\mathrm{\Omega }=\mathrm{d}\zeta \text{(on }W\text{)}$$ As above, changing $``$ by $`W`$, we obtain the Hamiltonian for $`𝒳`$, $`\{\zeta \}\mathrm{\Omega }^{km1}(W)/Z^{km1}(W)`$, and the local Hamiltonian forms for $`𝒳`$. Conversely, $`\zeta \mathrm{\Omega }^k()`$ (resp. $`\zeta \mathrm{\Omega }^k(W)`$) is said to be a Hamiltonian $`k`$-form (resp. a local Hamiltonian $`k`$-form) if there exists a multivector field $`𝒳\text{X}^{mk}()`$ (resp. $`𝒳\text{X}^{mk}(W)`$) such that (1) holds (resp. on $`W`$). In particular, when $`k=0`$, that is, if $`\zeta \mathrm{C}^{\mathrm{}}()`$), then the existence of Hamiltonian $`m`$-multivector fields for $`\zeta `$ is assured (see ). ### 2.3 Multimomentum bundles (See, for instance, ). Let $`\pi :EM`$ be the configuration bundle of a field theory, (with $`dimM=m`$, $`dimE=n+m`$). There are several multimomentum bundle structures associated with it. First we have $`\mathrm{\Lambda }_2^m\mathrm{T}^{}E`$, which is the bundle of $`m`$-forms on $`E`$ vanishing by the action of two $`\pi `$-vertical vector fields. Furthermore, if $`J^1\pi EM`$ denotes the first-order jet bundle over $`E`$, the set made of the affine maps from $`J^1\pi `$ to $`\mathrm{\Lambda }^m\mathrm{T}^{}M`$, denoted as $`\mathrm{Aff}(J^1\pi ,\mathrm{\Lambda }^m\mathrm{T}^{}M)`$, is another bundle over $`E`$ which is canonically diffeomorphic to $`\mathrm{\Lambda }_2^m\mathrm{T}^{}E`$ , . We will denote $$\pi \mathrm{\Lambda }_2^m\mathrm{T}^{}E\mathrm{Aff}(J^1\pi ,\mathrm{\Lambda }^m\mathrm{T}^{}M)$$ It is called the extended multimomentum bundle, and its canonical submersions are denoted $$\kappa :\pi E;\overline{\kappa }=\pi \kappa :\pi M$$ $`\pi `$ is a subbundle of $`\mathrm{\Lambda }^m\mathrm{T}^{}E`$, the multicotangent bundle of $`E`$ of order $`m`$ (the bundle of $`m`$-forms in $`E`$). Then $`\pi `$ is endowed with canonical forms. First we have the “tautological form” $`\mathrm{\Theta }\mathrm{\Omega }^m(\pi )`$ which is defined as follows: let $`(x,\alpha )\mathrm{\Lambda }_2^m\mathrm{T}^{}E`$, with $`xE`$ and $`\alpha \mathrm{\Lambda }_2^m\mathrm{T}_x^{}E`$; then, for every $`X_1,\mathrm{},X_m\mathrm{T}_{(x,\alpha )}(\pi )`$, $$\mathrm{\Theta }((x,\alpha );X_1,\mathrm{},X_m):=\alpha (x;\mathrm{T}_{(x,\alpha )}\kappa (X_1),\mathrm{},\mathrm{T}_{(q,\alpha )}\kappa (X_m))$$ Thus we define the multisymplectic form $$\mathrm{\Omega }:=\mathrm{d}\mathrm{\Theta }\mathrm{\Omega }^{m+1}(\pi )$$ They are known as the multimomentum Liouville $`m`$ and $`(m+1)`$-forms. We can introduce natural coordinates in $`\pi `$ adapted to the bundle $`\pi :EM`$, which are denoted by $`(x^\nu ,y^A,p_A^\nu ,p)`$, and such that $`\omega =\mathrm{d}^mx`$. Then the local expressions of these forms are $$\mathrm{\Theta }=p_A^\nu \mathrm{d}y^A\mathrm{d}^{m1}x_\nu +p\mathrm{d}^mx,\mathrm{\Omega }=\mathrm{d}p_A^\nu \mathrm{d}y^A\mathrm{d}^{m1}x_\nu \mathrm{d}p\mathrm{d}^mx$$ (2) (where $`d^{m1}x_\nu :=𝑖\left(\frac{}{x^\nu }\right)\mathrm{d}^mx`$). Consider $`\mathrm{\Lambda }_1^m\mathrm{T}^{}E\pi ^{}\mathrm{\Lambda }^m\mathrm{T}^{}M`$, which is another bundle over $`E`$, whose sections are the $`\pi `$-semibasic $`m`$-forms on $`E`$, and denote by $`J^1\pi ^{}`$ the quotient $`\mathrm{\Lambda }_2^m\mathrm{T}^{}E/\mathrm{\Lambda }_1^m\mathrm{T}^{}E\pi /\mathrm{\Lambda }_1^m\mathrm{T}^{}E`$. We have the natural submersions $$\tau :J^1\pi ^{}E;\overline{\tau }=\pi \tau :J^1\pi ^{}M$$ Furthermore, the natural submersion $`\mu :\pi J^1\pi ^{}`$ endows $`\pi `$ with the structure of an affine bundle over $`J^1\pi ^{}`$, with $`\tau ^{}\mathrm{\Lambda }_1^m\mathrm{T}^{}E`$ as the associated vector bundle. $`J^1\pi ^{}`$ is usually called the restricted multimomentum bundle associated with the bundle $`\pi :EM`$. Natural coordinates in $`J^1\pi ^{}`$ (adapted to the bundle $`\pi :EM`$) are denoted by $`(x^\nu ,y^A,p_A^\nu )`$. We have the diagram $$\begin{array}{ccc}\pi & \text{}& J^1\pi ^{}\\ & \text{}& \end{array}$$ Hamiltonian systems can be defined in $`\pi `$ or in $`J^1\pi ^{}`$. The construction of the Hamiltonian formalism in $`J^1\pi ^{}`$ was pionered in (see also and ), while a formulation in $`\pi `$ has been stated recently . In the following sections we review the main concepts of the formalism in $`J^1\pi ^{}`$, and we make an extensive development of the formalism in $`\pi `$. ## 3 Hamiltonian systems in $`J^1\pi ^{}`$ First we consider the standard definition of Hamiltonian systems in field theory, which is stated using the restricted multimomentum bundle $`J^1\pi ^{}`$. ### 3.1 Restricted Hamiltonian systems ###### Definition 1 Consider the bundle $`\overline{\tau }:J^1\pi ^{}M`$. 1. A section $`h:J^1\pi ^{}\pi `$ of the projection $`\mu `$ is called a Hamiltonian section of $`\mu `$. 2. The differentiable forms $$\mathrm{\Theta }_h:=h^{}\mathrm{\Theta },\mathrm{\Omega }_h:=\mathrm{d}\mathrm{\Theta }_h=h^{}\mathrm{\Omega }$$ are called the Hamilton-Cartan $`m`$ and $`(m+1)`$ forms of $`J^1\pi ^{}`$ associated with the Hamiltonian section $`h`$. 3. The couple $`(J^1\pi ^{},h)`$ is said to be a restricted Hamiltonian system, (or just a Hamiltonian system). In a local chart of natural coordinates, a Hamiltonian section is specified by a local Hamiltonian function $`\mathrm{h}\mathrm{C}^{\mathrm{}}(U)`$, $`UJ^1\pi ^{}`$, such that $`h(x^\nu ,y^A,p_A^\nu )(x^\nu ,y^A,p_A^\nu ,p=\mathrm{h}(x^\gamma ,y^B,p_B^\eta ))`$. The local expressions of the Hamilton-Cartan forms associated with $`h`$ are $$\mathrm{\Theta }_h=p_A^\nu \mathrm{d}y^A\mathrm{d}^{m1}x_\nu \mathrm{hd}^mx,\mathrm{\Omega }_h=\mathrm{d}p_A^\nu \mathrm{d}y^A\mathrm{d}^{m1}x_\nu +\mathrm{dh}\mathrm{d}^mx$$ (3) ###### Remark 1 Notice that $`\mathrm{\Omega }_h`$ is $`1`$-nondegenerate; that is, a multisymplectic form (as a simple calculation in coordinates shows). Hamiltonian sections can be obtained from connections. In fact, if we have a connection $``$ in $`\pi :EM`$, it induces a linear section $`h^{}:J^1\pi ^{}\pi `$ of $`\mu `$ . Then, if $`\mathrm{\Theta }`$ is the canonical $`m`$-form in $`\mathrm{\Omega }^m(\pi )`$, the forms $$\mathrm{\Theta }_h^{}:=h^{}\mathrm{\Theta }\mathrm{\Omega }^m(J^1\pi ^{}),\mathrm{\Omega }_h^{}:=\mathrm{d}\mathrm{\Theta }_h^{}\mathrm{\Omega }^{m+1}(J^1\pi ^{})$$ (4) are the Hamilton-Cartan $`m`$ and $`(m+1)`$ forms of $`J^1\pi ^{}`$ associated with the connection $``$. In a system of natural coordinates in $`J^1\pi ^{}`$, if $`=\mathrm{d}x^\nu \left({\displaystyle \frac{}{x^\nu }}+\mathrm{\Gamma }_\nu ^A{\displaystyle \frac{}{y^A}}\right)`$ is the local expression of the connection $``$, the local expressions of these Hamilton-Cartan forms associated with $``$ are $`\mathrm{\Theta }_h^{}`$ $`=`$ $`p_A^\nu (\mathrm{d}y^A\mathrm{\Gamma }_\eta ^A\mathrm{d}x^\eta )\mathrm{d}^{m1}x_\nu =p_A^\nu \mathrm{d}y^A\mathrm{d}^{m1}x_\nu p_A^\nu \mathrm{\Gamma }_\nu ^A\mathrm{d}^mx`$ $`\mathrm{\Omega }_h^{}`$ $`=`$ $`\mathrm{d}p_A^\nu \mathrm{d}y^A\mathrm{d}^{m1}x_\nu +\mathrm{\Gamma }_\nu ^A\mathrm{d}p_A^\nu \mathrm{d}^mx+p_A^\nu \mathrm{d}\mathrm{\Gamma }_\nu ^A\mathrm{d}^mx`$ Observe that a local Hamiltonian function associated with $`h^{}`$ is $`\mathrm{h}^{}=p_A^\nu \mathrm{\Gamma }_\nu ^A`$. ### 3.2 Variational principle and field equations Now we establish the field equations for restricted Hamiltonian systems. They can be derived from a variational principle. In fact, first we state: ###### Definition 2 Let $`(J^1\pi ^{},h)`$ be a restricted Hamiltonian system. Let $`\mathrm{\Gamma }(M,J^1\pi ^{})`$ be the set of sections of $`\overline{\tau }`$. Consider the map $$\begin{array}{ccccc}𝐇& :& \mathrm{\Gamma }(M,J^1\pi ^{})& & \\ & & \psi & & _M\psi ^{}\mathrm{\Theta }_h\end{array}$$ (where the convergence of the integral is assumed). The variational problem for this restricted Hamiltonian system is the search for the critical (or stationary) sections of the functional $`𝐇`$, with respect to the variations of $`\psi `$ given by $`\psi _t=\sigma _t\psi `$, where $`\{\sigma _t\}`$ is the local one-parameter group of any compact-supported $`Z\text{X}^{\mathrm{V}(\overline{\tau })}(J^1\pi ^{})`$ (where $`\text{X}^{\mathrm{V}(\overline{\tau })}(J^1\pi ^{})`$ denotes the module of $`\overline{\tau }`$-vertical vector fields in $`J^1\pi ^{}`$), that is: $$\frac{d}{dt}|_{t=0}_M\psi _t^{}\mathrm{\Theta }_h=0$$ This is the so-called Hamilton-Jacobi principle of the Hamiltonian formalism. Then the following fundamental theorem is proven (see also ): ###### Theorem 1 Let $`(J^1\pi ^{},h)`$ be a restricted Hamiltonian system. The following assertions on a section $`\psi \mathrm{\Gamma }(M,J^1\pi ^{})`$ are equivalent: 1. $`\psi `$ is a critical section for the variational problem posed by the Hamilton-Jacobi principle. 2. $`\psi ^{}𝑖(Z)\mathrm{\Omega }_h=0`$, for every $`Z\text{X}^{\mathrm{V}(\overline{\tau })}(J^1\pi ^{})`$. 3. $`\psi ^{}𝑖(X)\mathrm{\Omega }_h=0`$, for every $`X\text{X}(J^1\pi ^{})`$. 4. If $`(U;x^\nu ,y^A,p_A^\nu )`$ is a natural system of coordinates in $`J^1\pi ^{}`$, then $`\psi `$ satisfies the following system of equations in $`U`$ $$\frac{(y^A\psi )}{x^\nu }=\frac{\mathrm{h}}{p_A^\nu }\psi \frac{\mathrm{h}}{p_A^\nu }|_\psi ;\frac{(p_A^\nu \psi )}{x^\nu }=\frac{\mathrm{h}}{y^A}\psi \frac{\mathrm{h}}{y^A}|_\psi $$ (5) where $`\mathrm{h}`$ is a local Hamiltonian function associated with $`h`$. They are known as the Hamilton-De Donder-Weyl equations of the restricted Hamiltonian system. ( Proof ) ($`12`$) We assume that $`U`$ is a $`(m1)`$-dimensional manifold and that $`\overline{\tau }(supp(Z))U`$, for every compact-supported $`Z\text{X}^{\mathrm{V}(\overline{\tau })}(J^1\pi ^{})`$. Then $`{\displaystyle \frac{d}{dt}}|_{t=0}{\displaystyle _U}\psi _t^{}\mathrm{\Theta }_h`$ $`=`$ $`{\displaystyle \frac{d}{dt}}|_{t=0}{\displaystyle _U}\psi ^{}(\sigma _t^{}\mathrm{\Theta }_h)={\displaystyle _U}\psi ^{}\left(\underset{t0}{lim}{\displaystyle \frac{\sigma _t^{}\mathrm{\Theta }_h\mathrm{\Theta }_h}{t}}\right)`$ $`=`$ $`{\displaystyle _U}\psi ^{}(L(Z)\mathrm{\Theta }_h)={\displaystyle _U}\psi ^{}(𝑖(Z)\mathrm{d}\mathrm{\Theta }_h+\mathrm{d}𝑖(Z)\mathrm{\Theta }_h)`$ $`=`$ $`{\displaystyle _U}\psi ^{}(𝑖(Z)\mathrm{\Omega }_h\mathrm{d}𝑖(Z)\mathrm{\Theta }_h)={\displaystyle _U}\psi ^{}(𝑖(Z)\mathrm{\Omega }_h)+{\displaystyle _U}\mathrm{d}[\psi ^{}(𝑖(Z)\mathrm{\Theta }_h)]`$ $`=`$ $`{\displaystyle _U}\psi ^{}(𝑖(Z)\mathrm{\Omega }_h)+{\displaystyle _U}\psi ^{}(𝑖(Z)\mathrm{\Theta }_h)={\displaystyle _U}\psi ^{}(𝑖(Z)\mathrm{\Omega }_h)`$ (as a consequence of Stoke’s theorem and the hypothesis made on the supports of the vertical fields). Thus, by the fundamental theorem of the variational calculus we conclude that $`{\displaystyle \frac{d}{dt}}|_{t=0}{\displaystyle _U}\psi _t^{}\mathrm{\Theta }_h=0`$ $``$ $`\psi ^{}(𝑖(Z)\mathrm{\Omega }_h)=0`$, for every compact-supported $`Z\text{X}^{\mathrm{V}(\overline{\tau })}(J^1\pi ^{})`$. However, as compact-supported vector fields generate locally the $`\mathrm{C}^{\mathrm{}}(J^1\pi ^{})`$-module of vector fields in $`J^1\pi ^{}`$, it follows that the last equality holds for every $`Z\text{X}^{\mathrm{V}(\overline{\tau })}(J^1\pi ^{})`$. ($`23`$) If $`\mathrm{p}\mathrm{Im}\psi `$, then $`\mathrm{T}_\mathrm{p}J^1\pi ^{}=\mathrm{V}_\mathrm{p}(\overline{\tau })\mathrm{T}_\mathrm{p}(\mathrm{Im}\psi )`$. So if $`X\text{X}(J^1\pi ^{})`$, then $$X_\mathrm{p}=(X_\mathrm{p}\mathrm{T}_\mathrm{p}(\psi \overline{\tau })(X_\mathrm{p}))+\mathrm{T}_\mathrm{p}(\psi \overline{\tau })(X_\mathrm{p})X_\mathrm{p}^V+X_\mathrm{p}^\psi $$ and therefore $$\psi ^{}(𝑖(X)\mathrm{\Omega }_h)=\psi ^{}(𝑖(X^V)\mathrm{\Omega }_h)+\psi ^{}(𝑖(X^\psi )\mathrm{\Omega }_h)=\psi ^{}(𝑖(X^\psi )\mathrm{\Omega }_h)=0$$ since $`\psi ^{}(𝑖(X^V)\mathrm{\Omega }_h)=0`$ by the above item, and furthermore, $`X_\mathrm{p}^\psi \mathrm{T}_\mathrm{p}(\mathrm{Im}\psi )`$, and $`dim(\mathrm{Im}\psi )=m`$, being $`\mathrm{\Omega }_h\mathrm{\Omega }^{m+1}(J^1\pi ^{})`$. Hence we conclude that $`\psi ^{}(𝑖(X)\mathrm{\Omega }_h)=0`$, for every $`X\text{X}(J^1\pi ^{})`$. The converse is proved reversing this reasoning. ($`34`$) If $`X=\alpha ^\nu {\displaystyle \frac{}{x^\nu }}+\beta ^A{\displaystyle \frac{}{y^A}}+\gamma _A^\nu {\displaystyle \frac{}{p_A^\nu }}\text{X}(J^1\pi ^{})`$ , taking into account the local expression (3) of $`\mathrm{\Omega }_h`$, we have $`𝑖(X)\mathrm{\Omega }_h`$ $`=`$ $`(1)^\eta \alpha ^\eta \left(\mathrm{d}p_A^\nu \mathrm{d}y^A\mathrm{d}^{m2}x_{\eta \nu }{\displaystyle \frac{\mathrm{h}}{p_A^\nu }}\mathrm{d}p_A^\nu \mathrm{d}^{m1}x_\eta \right)`$ $`+`$ $`\beta ^A\left(\mathrm{d}p_A^\nu \mathrm{d}^{m1}x_\nu +{\displaystyle \frac{\mathrm{h}}{y^A}}\mathrm{d}^mx\right)+\gamma _A^\nu \left(\mathrm{d}y^A\mathrm{d}^{m1}x_\nu +{\displaystyle \frac{\mathrm{h}}{p_A^\nu }}\mathrm{d}^mx\right)`$ but if $`\psi =(x^\nu ,y^A(x^\eta ),p_A^\nu (x^\eta ))`$, then $`\psi ^{}𝑖(X)\mathrm{\Omega }_h`$ $`=`$ $`(1)^{\eta +\nu }\alpha ^\eta \left({\displaystyle \frac{(y^A\psi )}{x^\nu }}{\displaystyle \frac{\mathrm{h}}{p_A^\nu }}|_\psi \right){\displaystyle \frac{(p_A^\nu \psi )}{x^\eta }}\mathrm{d}^mx`$ $`+\beta ^A\left({\displaystyle \frac{(p_A^\nu \psi )}{x^\nu }}+{\displaystyle \frac{\mathrm{h}}{y^A}}|_\psi \right)\mathrm{d}^mx+\gamma _A^\nu \left({\displaystyle \frac{(y^A\psi )}{x^\nu }}+{\displaystyle \frac{\mathrm{h}}{p_A^\nu }}|_\psi \right)\mathrm{d}^mx`$ and, as this holds for every $`X\text{X}(J^1\pi ^{})`$, we conclude that $`\psi ^{}𝑖(X)\mathrm{\Omega }_h=0`$ if, and only if, the Hamilton-De Donder-Weyl equations hold for $`\psi `$. ###### Remark 2 It is important to point out that equations (5) are not covariant, since the Hamiltonian function $`\mathrm{h}`$ is defined only locally, and hence it is not intrinsically defined. In order to write a set of covariant Hamiltonian equations we must use a global Hamiltonian function, that is, a Hamiltonian density (see and for comments on this subject). Observe also that the solution to these equations is not unique. ### 3.3 Hamiltonian equations for multivector fields (See and for more details). Let $`(J^1\pi ^{},h)`$ be a restricted Hamiltonian system. The problem of finding critical sections solutions of the Hamilton-Jacobi principle can be formulated equivalently as follows: to find a distribution $`D`$ of $`\mathrm{T}(J^1\pi ^{})`$ satisfying that: * $`D`$ is $`m`$-dimensional. * $`D`$ is $`\overline{\tau }`$-transverse. * $`D`$ is integrable (that is, involutive). * The integral manifolds of $`D`$ are the critical sections of the Hamilton-Jacobi principle. However, as explained in Section 2.1, these kinds of distributions are associated with classes of integrable (i.e., non-vanishing, locally decomposable and involutive) $`\overline{\tau }`$-transverse multivector fields in $`J^1\pi ^{}`$. The local expression in natural coordinates of an element of one of these classes is $$𝒳=\underset{\nu =1}{\overset{m}{}}f\left(\frac{}{x^\nu }+F_\nu ^A\frac{}{y^A}+G_{A\nu }^\rho \frac{}{p_A^\rho }\right)$$ (6) where $`f\mathrm{C}^{\mathrm{}}(J^1\pi ^{})`$ is a non-vanishing function. Therefore, the problem posed by the Hamilton-Jacobi principle can be stated in the following way , : ###### Theorem 2 Let $`(J^1\pi ^{},h)`$ be a restricted Hamiltonian system, and $`\{𝒳\}\text{X}^m(J^1\pi ^{})`$ a class of integrable and $`\overline{\tau }`$-transverse multivector fields. Then, the integral manifolds of $`\{𝒳\}`$ are critical section for the variational problem posed by the Hamilton-Jacobi principle if, and only if, $$𝑖(𝒳_h)\mathrm{\Omega }_h=0,\text{for every }𝒳_h\{𝒳_h\}$$ (7) ###### Remark 3 The $`\overline{\tau }`$-transversality condition for multivector fields solution to (7) can be stated by demanding that $`𝑖(𝒳_h)(\overline{\tau }^{}\omega )0`$. In particular, if we take $`𝑖(𝒳_h)(\overline{\tau }^{}\omega )=1`$ we are choosing a representative of the class of $`\overline{\tau }`$-transverse multivector fields solution to (7). (This is equivalent to putting $`f=1`$ in the local expression (6)). Thus, the problem posed in Definition 2 is equivalent to looking for a multivector field $`𝒳_h\text{X}^m(J^1\pi ^{})`$ such that: 1. $`𝑖(𝒳_h)\mathrm{\Omega }_h=0`$. 2. $`𝑖(𝒳_h)(\overline{\tau }^{}\omega )=1`$. 3. $`𝒳_h`$ is integrable. From the conditions 1 and 2, using the local expressions (3) of $`\mathrm{\Omega }_h`$ and (6) for $`𝒳_h`$, we obtain that $`f=1`$ and $$F_\nu ^A=\frac{\mathrm{h}}{p_A^\nu };G_{A\nu }^\nu =\frac{\mathrm{h}}{y^A}$$ and, if $`\psi (x)=(x^\nu ,y^A(x^\gamma ),p_A^\nu (x^\gamma ))`$ must be an integral section of $`𝒳_h`$, then $$\frac{(y^A\psi )}{x^\nu }=F_\nu ^A\psi ;\frac{(p_A^\rho \psi )}{x^\nu }=G_{A\nu }^\rho \psi $$ Thus the Hamilton-De Donder-Weyl equations (5) for $`\psi `$ are recovered from (7). ###### Remark 4 Classes of locally decomposable and $`\overline{\tau }`$-transverse multivector fields are in one-to one correspondence with connections in the bundle $`\overline{\tau }:J^1\pi ^{}M`$ (see Section 2.1). Then, it can be proven that the condition stated in Theorem 7 is equivalent to finding an integrable connection $`_h`$ in $`J^1\pi ^{}M`$ satisfying the equation $$𝑖(_h)\mathrm{\Omega }_h=(m1)\mathrm{\Omega }_h$$ whose integral sections are the critical sections of the Hamilton-Jacobi problem. Of course, $`_h`$ is the connection associated to the class $`\{𝒳_h\}`$ solution to (7), and $`𝒳_h`$ is integrable if, and only if, the curvature of $`_h`$ vanishes everywhere. The expression of $`_h`$ in coordinates is $$_h=\mathrm{d}x^\nu \left(\frac{}{x^\nu }+F_\nu ^A\frac{}{y^A}+G_{A\nu }^\rho \frac{}{p_A^\rho }\right)$$ ###### Definition 3 $`𝒳_h\text{X}^m(J^1\pi ^{})`$ will be called a Hamilton-De Donder-Weyl (HDW) multivector field for the system $`(J^1\pi ^{},h)`$ if it is $`\overline{\tau }`$-transverse, locally decomposable and verifies the equation $`𝑖(𝒳_h)\mathrm{\Omega }_h=0`$. Then, the associated connection $`_h`$ is called a Hamilton-De Donder-Weyl (HDW) connection for $`(J^1\pi ^{},h)`$. For restricted Hamiltonian systems, the existence of Hamilton-De Donder Weyl multivector fields or connections is guaranteed, although they are not necessarily integrable , : ###### Theorem 3 (Existence and local multiplicity of HDW-multivector fields): Let $`(J^1\pi ^{},h)`$ be a restricted Hamiltonian system. Then there exist classes of HDW-multivector fields $`\{𝒳_h\}`$. In a local system the above solutions depend on $`n(m^21)`$ arbitrary functions. ###### Remark 5 In order to find a class of integrable HDW-multivector fields (if it exists) we must impose that $`𝒳_h`$ verify the integrability condition: the curvature of $`_h`$ vanishes everywhere. Hence the number of arbitrary functions will in general be less than $`n(m^21)`$. If this integrable multivector field does not exist, we can eventually select some particular HDW-multivector field solution, and apply an integrability algorithm in order to find a submanifold $`J^1\pi ^{}`$ (if it exists), where this multivector field is integrable (and tangent to $``$). ## 4 Hamiltonian systems in $`\pi `$ Now we introduce Hamiltonian systems in the extended multimomentum bundle $`\pi `$, and we study their relation with those defined in the above section. ### 4.1 Extended Hamiltonian systems Now we have the multisymplectic manifold $`(\pi ,\mathrm{\Omega })`$, and we are interested in defining Hamiltonian systems on this manifold which are suitable for describing Hamiltonian field theories. Thus we must consider Hamiltonian or locally Hamiltonian $`m`$-multivector fields and forms of a particular kind. In particular, bearing in mind the requirements in Remark 3, we can state: ###### Definition 4 The triple $`(\pi ,\mathrm{\Omega },\alpha )`$ is said to be an extended Hamiltonian system if: 1. $`\alpha Z^1(\pi )`$ (it is a closed $`1`$-form). 2. There exists a locally decomposable multivector field $`𝒳_\alpha \text{X}^m(\pi )`$ satisfying that $$𝑖(𝒳_\alpha )\mathrm{\Omega }=(1)^{m+1}\alpha ,𝑖(𝒳_\alpha )(\overline{\kappa }^{}\omega )=1\text{(}\overline{\kappa }\text{-transversality)}$$ (8) If $`\alpha `$ is an exact form, then $`(\pi ,\mathrm{\Omega },\alpha )`$ is an extended global Hamiltonian system. In this case, there exist $`\mathrm{H}\mathrm{C}^{\mathrm{}}(\pi )`$ such that $`\alpha =\mathrm{dH}`$, which are called Hamiltonian functions of the system. (For an extended Hamiltonian system, these functions exist only locally, and they are called local Hamiltonian functions). The condition that $`\alpha `$ is closed plays a crucial role (see Proposition 2 and Section 4.2). The factor $`(1)^{m+1}`$ in the definition will be justified later (see Proposition 10 and Remark 7). Observe that, if $`(\pi ,\mathrm{\Omega },\alpha )`$ is an extended global Hamiltonian system, giving a Hamiltonian function $`\mathrm{H}`$ is equivalent to giving a Hamiltonian density $`\stackrel{~}{}\mathrm{H}(\overline{\kappa }^{}\omega )\mathrm{\Omega }^m(\pi )`$. In natural coordinates of $`\pi `$, the most general expression for a locally decomposable multivector field $`𝒳_\alpha \text{X}^m(\pi `$ is $$𝒳_\alpha =\underset{\nu =1}{\overset{m}{}}\stackrel{~}{f}\left(\frac{}{x^\nu }+\stackrel{~}{F}_\nu ^A\frac{}{y^A}+\stackrel{~}{G}_{A\nu }^\rho \frac{}{p_A^\rho }+\stackrel{~}{g}_\nu \frac{}{p}\right)$$ (9) where $`\stackrel{~}{f}\mathrm{C}^{\mathrm{}}(\pi )`$ is a non-vanishing function which is equal to $`1`$ if the equation $`𝑖(𝒳_\alpha )(\overline{\kappa }^{}\omega )=1`$ holds. ###### Remark 6 In addition, bearing in mind Remark 5, the integrability of $`𝒳_\alpha `$ must be imposed. Then all the multivector fields in the integrable class $`\{𝒳_\alpha \}`$ have the same integral sections. A first important observation is that not every closed form $`\alpha \mathrm{\Omega }^m(\pi )`$ defines an extended Hamiltonian system. In fact: ###### Proposition 1 If $`(\pi ,\mathrm{\Omega },\alpha )`$ is an extended Hamiltonian system, then $`𝑖(Y)\alpha 0`$, for every $`\mu `$-vertical vector field $`Y\text{X}^{\mathrm{V}(\mu )}(\pi )`$, $`Y0`$. In particular, for every system of natural coordinates $`(x^\nu ,y^A,p_A^\nu ,p))`$ in $`\pi `$ adapted to the bundle $`\pi :EM`$ (with $`\omega =\mathrm{d}^mx`$), $$𝑖\left(\frac{}{p}\right)\alpha =1$$ (10) ( Proof ) In order to prove this, we use natural coordinates of $`\pi `$. The local expression of $`\mathrm{\Omega }`$ is given in (2), and a $`\mu `$-vertical vector field is locally given by $`Y=f{\displaystyle \frac{}{p}}`$. Then, if $`𝒳_\alpha \text{X}^m(\pi )`$ is a multivector field solution to the equations (8), we have $`𝑖(Y)\alpha `$ $`=`$ $`(1)^{m+1}𝑖(Y)𝑖(𝒳_\alpha )\mathrm{\Omega }=(1)^{m+1}(1)^m𝑖(𝒳_\alpha )𝑖(Y)\mathrm{\Omega }`$ $`=`$ $`𝑖(𝒳_\alpha )𝑖\left(f{\displaystyle \frac{}{p}}\right)[\mathrm{d}p\mathrm{d}^mx\mathrm{d}p_A^\nu \mathrm{d}y^A\mathrm{d}^{m1}x_\nu ]`$ $`=`$ $`𝑖(𝒳_\alpha )𝑖\left(f{\displaystyle \frac{}{p}}\right)[\mathrm{d}p\mathrm{d}^mx]=f𝑖(𝒳_\alpha )\mathrm{d}^mx=f`$ and, as $`Y0f0`$, the first result holds. In particular, taking $`f=1`$, the expression (10) is reached. As a consequence of this result we have: ###### Proposition 2 If $`(\pi ,\mathrm{\Omega },\alpha )`$ is an extended Hamiltonian system, locally $`\alpha =\mathrm{d}p+\beta `$, where $`\beta `$ is a closed and $`\mu `$-basic local $`1`$-form in $`\pi `$. ( Proof ) As a consequence of (10), $`\alpha =\mathrm{d}p+\beta `$ locally, where $`\beta `$ is a $`\mu `$-semibasic local $`1`$-form. But, as $`\alpha `$ is closed, so is $`\beta `$. Hence, for every $`Y\text{X}^{\mathrm{V}(\mu )}(\pi )`$, we have that $`L(Y)\beta =𝑖(Y)\mathrm{d}\beta +\mathrm{d}𝑖(Y)\beta =0`$, and $`\beta `$ is $`\mu `$-basic. Therefore, by Poincaré’s lemma, on an open set $`U\pi `$ $`\alpha `$ has necessarily the following coordinate expression $$\alpha =\mathrm{d}p+\mathrm{d}\stackrel{~}{\mathrm{h}}(x^\nu ,y^A,p_A^\nu )$$ (11) where $`\stackrel{~}{\mathrm{h}}=\mu ^{}\mathrm{h}`$, for some $`\mathrm{h}\mathrm{C}^{\mathrm{}}(\mu (U))`$. Then, if $`\mathrm{H}`$ is a (local) Hamiltonian function for $`\alpha `$; that is, such that $`\alpha =\mathrm{dH}`$ (at least locally), we have that (see also ) $$\mathrm{H}=p+\stackrel{~}{\mathrm{h}}(x^\nu ,y^A,p_A^\nu )$$ (12) where $`\stackrel{~}{\mathrm{h}}(x^\nu ,y^A,p_A^\nu )`$ is determined up to a constant. Conversely, every closed form $`\alpha \mathrm{\Omega }^1(\pi )`$ satisfying the above condition defines an extended Hamiltonian system since, in an analogous way to Theorem 3, we can prove: ###### Theorem 4 Let $`\alpha Z^1(\pi )`$ satisfying the condition stated in Propositions 10 and 2. Then there exist locally decomposable multivector fields $`𝒳_\alpha \text{X}^m(\pi )`$ (not necessarily integrable) satisfying equations (8) (and hence $`(\pi ,\mathrm{\Omega },\alpha )`$ is an extended Hamiltonian system). In a local system the above solutions depend on $`n(m^21)`$ arbitrary functions. ( Proof ) We use the local expressions (2), (11) and (9) for $`\mathrm{\Omega }`$, $`\alpha `$ and $`𝒳_\alpha `$ respectively. Then $`𝑖(𝒳_\alpha )(\overline{\kappa }^{}\omega )=1`$ leads to $`\stackrel{~}{f}=1`$. Furthermore, from $`𝑖(𝒳_\alpha )\mathrm{\Omega }=(1)^{m+1}\alpha `$ we obtain that the equality for the coefficients on $`\mathrm{d}p_A^\nu `$ leads to $$\stackrel{~}{F}_\nu ^A=\frac{\mathrm{H}}{p_A^\nu }=\frac{\stackrel{~}{\mathrm{h}}}{p_A^\nu }(\text{for every }A,\nu )$$ (13) For the coefficients on $`\mathrm{d}y^A`$ we have $$\stackrel{~}{G}_{A\nu }^\nu =\frac{\mathrm{H}}{y^A}=\frac{\stackrel{~}{\mathrm{h}}}{y^A}(A=1,\mathrm{},n)$$ (14) and for the coefficients on $`\mathrm{d}x^\nu `$, using these results, we obtain $`\stackrel{~}{g}_\nu `$ $`=`$ $`{\displaystyle \frac{\mathrm{H}}{x^\nu }}+\stackrel{~}{F}_\nu ^A\stackrel{~}{G}_{A\eta }^\eta \stackrel{~}{F}_\eta ^A\stackrel{~}{G}_{A\nu }^\eta ={\displaystyle \frac{\mathrm{H}}{x^\nu }}+{\displaystyle \frac{\mathrm{H}}{p_A^\nu }}\stackrel{~}{G}_{A\eta }^\eta {\displaystyle \frac{\mathrm{H}}{p_A^\eta }}\stackrel{~}{G}_{A\nu }^\eta `$ (15) $`=`$ $`{\displaystyle \frac{\stackrel{~}{\mathrm{h}}}{x^\nu }}+{\displaystyle \frac{\stackrel{~}{\mathrm{h}}}{p_A^\nu }}\stackrel{~}{G}_{A\eta }^\eta {\displaystyle \frac{\stackrel{~}{\mathrm{h}}}{p_A^\eta }}\stackrel{~}{G}_{A\nu }^\eta (A=1,\mathrm{},n;\eta \nu )`$ where the coefficients $`G_{A\nu }^\rho `$ are related by the equations (14). Finally, the coefficient on $`\mathrm{d}p`$ are identical, taking into account the above results. Thus, equations (13) make a system of $`nm`$ linear equations which determines univocally the functions $`\stackrel{~}{F}_\nu ^A`$, equations (14) are a compatible system of $`n`$ linear equations on the $`nm^2`$ functions $`\stackrel{~}{G}_{A\nu }^\gamma `$, and equations (15) make a system of $`m`$ linear equations which determines univocally the functions $`\stackrel{~}{g}_\nu `$. In this way, solutions to equations (8) are determined locally from the relations (13) and (15), and through the $`n`$ independent linear equations (14). Therefore, there are $`n(m^21)`$ arbitrary functions. These results assure the local existence of $`𝒳_\alpha `$. The global solutions are obtained using a partition of unity subordinated to a covering of $`\pi `$ made of natural charts. (A further local analysis of these multivector fields solution and other additional details can be found in and ). ###### Remark 7 With regard to this result, it is important to point out that, if $`𝒳_\alpha \text{X}^m(\pi )`$ is a solution (not necessarily integrable) to the equations (8), then every multivector field $`𝒳_\alpha ^{}\{𝒳_\alpha \}`$; that is, such that $`𝒳_\alpha ^{}=\stackrel{~}{f}𝒳_\alpha `$ (where $`\stackrel{~}{f}\mathrm{C}^{\mathrm{}}(\pi )`$ is non-vanishing) is a solution to the equations $$𝑖(𝒳_\alpha ^{})\mathrm{\Omega }=\stackrel{~}{f}(1)^{m+1}\alpha ,𝑖(𝒳_\alpha ^{})(\overline{\kappa }^{}\omega )=\stackrel{~}{f}\text{(}\overline{\kappa }\text{-transversality)}$$ In particular, if we have a $`1`$-form $`\alpha =\mathrm{dH}`$ (locally), with $`0{\displaystyle \frac{\mathrm{H}}{p}}1`$, then the $`\overline{\kappa }`$-transversality condition must be stated as $`𝑖(𝒳_\alpha )(\overline{\kappa }^{}\omega )={\displaystyle \frac{\mathrm{H}}{p}}`$, and the solutions $`𝒳_\alpha `$ to the equation $`𝑖(𝒳_\alpha )\mathrm{\Omega }=(1)^{m+1}\alpha `$ have the local expression (9) with $`\stackrel{~}{f}={\displaystyle \frac{\mathrm{H}}{p}}`$, and the other coefficients being solutions to the system of equations $$\stackrel{~}{f}\stackrel{~}{F}_\nu ^A=\frac{\mathrm{H}}{p_A^\nu },\stackrel{~}{f}\stackrel{~}{G}_{A\nu }^\nu =\frac{\mathrm{H}}{y^A},\stackrel{~}{f}\stackrel{~}{g}_\nu =\frac{\mathrm{H}}{x^\nu }+\frac{\mathrm{H}}{p_A^\nu }\stackrel{~}{G}_{A\eta }^\eta \frac{\mathrm{H}}{p_A^\eta }\stackrel{~}{G}_{A\nu }^\eta (\eta \nu )$$ Thus, in an analogous way to restricted Hamiltonian systems in $`J^1\pi ^{}`$, we define: ###### Definition 5 $`𝒳_\alpha \text{X}^m(\pi )`$ will be called an extended Hamilton-De Donder-Weyl multivector field for the system $`(\pi ,\mathrm{\Omega },\alpha )`$ if it is $`\overline{\kappa }`$-transverse, locally decomposable and verifies the equation $`𝑖(𝒳_\alpha )\mathrm{\Omega }=(1)^{m+1}\alpha `$. Then, the associated connection $`_\alpha `$ in the bundle $`\overline{\kappa }:\pi M`$ is called an extended Hamilton-De Donder-Weyl connection for $`(\pi ,\mathrm{\Omega },\alpha )`$. Now, if $`\{X_\alpha \}`$ is integrable and $`\stackrel{~}{\psi }(x)=(x^\nu ,y^A(x^\gamma ),p_A^\nu (x^\gamma ),p(x^\gamma ))`$ must be an integral section of $`𝒳_\alpha `$ then $$\frac{(y^A\stackrel{~}{\psi })}{x^\nu }=\stackrel{~}{F}_\nu ^A\stackrel{~}{\psi },\frac{(p_A^\rho \stackrel{~}{\psi })}{x^\nu }=\stackrel{~}{G}_{A\nu }^\rho \stackrel{~}{\psi },\frac{(p\stackrel{~}{\psi })}{x^\nu }=\stackrel{~}{g}_\nu \stackrel{~}{\psi }$$ (16) so equations (13), (14) and (15) give PDE’s for $`\stackrel{~}{\psi }`$. In particular, the Hamilton-De Donder-Weyl equations (5) are recovered from (13) and (14). As for restricted Hamiltonian systems, in order to find a class of integrable extended HDW-multivector fields (if it exists) we must impose that $`𝒳_\alpha `$ verify the integrability condition, that is, that the curvature of $`_\alpha `$ vanishes everywhere, and thus the number of arbitrary functions will in general be less than $`n(m^21)`$. Just as in that situation, we cannot assure the existence of an integrable solution. If it does not exist, we can eventually select some particular extended HDW-multivector field solution, and apply an integrability algorithm in order to find a submanifold of $`\pi `$ (if it exists), where this multivector field is integrable (and tangent to it). ### 4.2 Geometric properties of extended Hamiltonian systems Most of the properties of the extended Hamiltonian systems are based in the following general results: ###### Lemma 1 Let $`\mu :`$ be a surjective submersion, with $`dim=dim+r`$. Consider $`\alpha _1,\mathrm{},\alpha _r\Omega ^1()`$ such that $`\alpha \alpha _1\mathrm{}\alpha _r`$ is a closed $`r`$-form, and $`\alpha (𝐩)0`$, for every $`𝐩`$. Finally, let $`\{\alpha \}^0:=\{Z\text{X}()|𝑖(Z)\alpha =0\}`$ be the annihilator of $`\alpha `$. Therefore: 1. $`\{\alpha \}^0:=\{Z\text{X}()|𝑖(Z)\alpha _i=0,i=1\mathrm{}r\}`$. 2. $`\{\alpha \}^0`$ generates an involutive distribution in $``$ of corank equal to $`r`$, which is called the characteristic distribution of $`\alpha `$, and is denoted $`𝒟_\alpha `$. If, in addition, the condition $`𝑖(Y)\alpha 0`$ holds, for every $`Y\text{X}^{\mathrm{V}(\mu )}()`$, then: 3. $`𝒟_\alpha `$ is a $`\mu `$-transverse distribution. 4. The integral submanifolds of $`𝒟_\alpha `$ are $`r`$-codimensional and $`\mu `$-transverse local submanifolds of $``$. 5. $`\mathrm{T}_𝐩=\mathrm{V}_𝐩(\mu )\{\alpha \}_𝐩^0`$, for every $`𝐩`$. 6. If $`S`$ is an integral submanifold of $`𝒟_\alpha `$, then $`\mu |_S:S`$ is a local diffeomorphism. 7. For every integral submanifold $`S`$ of $`𝒟_\alpha `$, and $`𝐩S`$, there exists $`W`$, with $`𝐩W`$, such that $`h=(\mu |_{WS})^1`$ is a local section of $`\mu `$ defined on $`\mu (WS)`$ (which is an open set of $``$). ( Proof ) First, observe that, for every $`𝐩`$, $`\alpha (𝐩)0`$ implies that $`\alpha _i(𝐩)`$, for every $`i=1,\mathrm{},r`$, are linearly independent, then $$0=𝑖(Z)\alpha =\underset{i=1}{\overset{r}{}}(1)^{i1}𝑖(Z)\alpha _i(\alpha _1\mathrm{}\alpha _{i1}\alpha _{i+1}\mathrm{}\alpha _r)𝑖(Z)\alpha _i=0$$ hence the statement in the item 1 holds and, as a consequence, we conclude that $`\{\alpha \}^0`$ generates a distribution in $``$ of rank equal to $`dim`$. Furthermore, if $`\alpha `$ is closed, for every $`Z_1,Z_2\{\alpha \}^0`$, we obtain that $`[Z_1,Z_2]\{\alpha \}^0`$ because $$𝑖([Z_1,Z_2])\alpha =L(Z_1)𝑖(Z_2)\alpha 𝑖(Z_2)L(Z_1)\alpha =𝑖(Z_2)[𝑖(Z_1)\mathrm{d}\alpha \mathrm{d}𝑖(Z_1)\alpha ]=0$$ Then $`𝒟_\alpha `$ is involutive. The other properties follow straighforwardly from these results and the condition $`𝑖(Y)\alpha 0`$, for every $`Y\text{X}^{\mathrm{V}(\mu )}()`$. Now, from this lemma we have that: ###### Proposition 3 If $`(\pi ,\mathrm{\Omega },\alpha )`$ is an extended Hamiltonian system, then: 1. $`𝒟_\alpha `$ is a $`\mu `$-transverse involutive distribution of corank equal to $`1`$. 2. The integral submanifolds $`S`$ of $`𝒟_\alpha `$ are $`1`$-codimensional and $`\mu `$-transverse local submanifolds of $`\pi `$. (We denote by $`ȷ_S:S\pi `$ the natural embedding). 3. For every $`𝐩\pi `$, we have that $`\mathrm{T}_𝐩\pi =\mathrm{V}_𝐩(\mu )(𝒟_\alpha )_𝐩`$, and thus, in this way, $`\alpha `$ defines an integrable connection in the affine bundle $`\mu :\pi J^1\pi ^{}`$. 4. If $`S`$ is an integral submanifold of $`𝒟_\alpha `$, then $`\mu |_S:SJ^1\pi ^{}`$ is a local diffeomorphism. 5. For every integral submanifold $`S`$ of $`𝒟_\alpha `$, and $`𝐩S`$, there exists $`W\pi `$, with $`𝐩W`$, such that $`h=(\mu |_{WS})^1`$ is a local Hamiltonian section of $`\mu `$ defined on $`\mu (WS)`$. ###### Remark 8 Observe that, if $`(\pi ,\mathrm{\Omega },\alpha )`$ is an extended Hamiltonian system, as $`\alpha =\mathrm{dH}`$ (locally), every local Hamiltonian function $`\mathrm{H}`$ is a constraint defining locally the integral submanifolds of $`𝒟_\alpha `$. Thus, bearing in mind the coordinate expression (12), the local Hamiltonian sections associated with these submanifolds are locally expressed as $$h(x^\nu ,y^A,p_A^\nu )=(x^\nu ,y^A,p_A^\nu ,p=\mathrm{h}(x^\gamma ,y^B,p_B^\eta ))$$ where $`\mu ^{}\mathrm{h}=\stackrel{~}{\mathrm{h}}`$. A relevant question is under what conditions the existence of global Hamiltonian sections is assured. The answer is given by the following: ###### Proposition 4 Let $`(\pi ,\mathrm{\Omega },\alpha )`$ be an extended global Hamiltonian system. If there is a Hamiltonian function $`\mathrm{H}\mathrm{C}^{\mathrm{}}(\pi )`$, and $`k`$, such that $`\mu (\mathrm{H}^1(k))=J^1\pi ^{}`$, then there exists a global Hamiltonian section $`h\mathrm{\Gamma }(J^1\pi ^{},\pi )`$. ( Proof ) In order to construct $`h`$, we prove that, for every $`\mathrm{q}J^1\pi ^{}`$, we have that $`\mu ^1(\mathrm{q})S_k`$ contains only one point. Let $`(U;x^\nu ,y^A,p_A^\nu ,p)`$ be a local chart in $`\pi `$, with $`\mathrm{q}\mu (U)`$. By Proposition 2 we have that $`\mathrm{H}|_U=p+\mu ^{}\mathrm{h}`$, for $`\mathrm{h}\mathrm{C}^{\mathrm{}}(\mu (U))`$. If $`𝐩\mu ^1(\mathrm{q})S_k`$ we have that $$k=\mathrm{H}(𝐩)=p(𝐩)(\mu ^{}\mathrm{h})(𝐩=p(𝐩)\mathrm{h}(\mu (𝐩))$$ then $`p(𝐩)`$ is determined by $`\mathrm{q}`$, and $`𝐩`$ is unique. This allows to define a global section $`h:J^1\pi ^{}\pi `$ by $`h(\mathrm{q}):=\mu ^1(\mathrm{q})S_k`$, for every $`\mathrm{q}J^1\pi ^{}`$, which obviously does not depend on the local charts considered. Observe that, if the first de Rahm cohomology group $`H^1(\pi )=0`$, then every extended Hamiltonian system is a global one, but this does not assure the existence of global Hamiltonian sections, as we have shown. In addition, we have: ###### Proposition 5 Given an extended Hamiltonian system $`(\pi ,\mathrm{\Omega },\alpha )`$, every extended HDW multivector field $`𝒳_\alpha \text{X}^m(\pi )`$ for the system $`(\pi ,\mathrm{\Omega },\alpha )`$ is tangent to every integral submanifold of $`𝒟_\alpha `$. As a consequence, if $`𝒳_\alpha `$ is integrable, then its integral sections are contained in the integral submanifolds of $`𝒟_\alpha `$. ( Proof ) By definition, an extended HDW multivector field is locally decomposable, so locally $`𝒳_\alpha =X_1\mathrm{}X_m`$, with $`X_1,\mathrm{},X_m\text{X}(\pi )`$. Then $`𝒳_\alpha `$ is tangent to every integral submanifold $`S`$ of $`𝒟_\alpha `$ if, and only if, $`X_\nu `$ are tangent to $`S`$, for every $`\nu =1,\mathrm{},m`$. But, as $`𝒟_\alpha `$ is the characteristic distribution of $`\alpha `$, this is equivalent to $`ȷ_S^{}𝑖(X_\nu )\alpha =0`$, and this is true because $$𝑖(X_\nu )\alpha =𝑖(X_\nu )𝑖(𝒳_\alpha )\mathrm{\Omega }=𝑖(X_\nu )𝑖(X_1\mathrm{}X_m)\mathrm{\Omega }=0$$ The last consequence is immediate. ###### Remark 9 Observe that, if $`𝒳_\alpha =X_1\mathrm{}X_m`$ locally, using the local expressions (9) and (11) and equations (13) and (14), the conditions $`𝑖(X_\nu )\alpha =0`$ lead to $`0`$ $`=`$ $`{\displaystyle \frac{\stackrel{~}{\mathrm{h}}}{x^\nu }}+\stackrel{~}{F}_\nu ^A{\displaystyle \frac{\stackrel{~}{\mathrm{h}}}{y^A}}+\stackrel{~}{G}_{A\nu }^\rho {\displaystyle \frac{\stackrel{~}{\mathrm{h}}}{p_A^\rho }}+\stackrel{~}{g}_\nu `$ $`=`$ $`{\displaystyle \frac{\stackrel{~}{\mathrm{h}}}{x^\nu }}{\displaystyle \frac{\stackrel{~}{\mathrm{h}}}{p_A^\rho }}\stackrel{~}{G}_{A\rho }^\rho +\stackrel{~}{G}_{A\nu }^\rho {\displaystyle \frac{\stackrel{~}{\mathrm{h}}}{p_A^\rho }}+\stackrel{~}{g}_\nu `$ $`=`$ $`{\displaystyle \frac{\stackrel{~}{\mathrm{h}}}{x^\nu }}{\displaystyle \frac{\stackrel{~}{\mathrm{h}}}{p_A^\rho }}(\stackrel{~}{G}_{A\eta }^\eta +\stackrel{~}{G}_{A\nu }^\nu )+\stackrel{~}{G}_{A\nu }^\eta {\displaystyle \frac{\stackrel{~}{\mathrm{h}}}{p_A^\eta }}+\stackrel{~}{G}_{A\nu }^\nu {\displaystyle \frac{\stackrel{~}{\mathrm{h}}}{p_A^\nu }}+\stackrel{~}{g}_\nu `$ $`=`$ $`{\displaystyle \frac{\stackrel{~}{\mathrm{h}}}{x^\nu }}{\displaystyle \frac{\stackrel{~}{\mathrm{h}}}{p_A^\nu }}\stackrel{~}{G}_{A\eta }^\eta +{\displaystyle \frac{\stackrel{~}{\mathrm{h}}}{p_A^\eta }}\stackrel{~}{G}_{A\nu }^\eta +\stackrel{~}{g}_\nu (\rho =1,\mathrm{},m;\nu \mathrm{fixed},\eta \nu )`$ which are the equations (15). So these equations are consistency conditions. (See also the comment in Remark 10). Finally, we have the following result: ###### Proposition 6 The integral submanifolds of $`𝒟_\alpha `$ are $`m`$-coisotropic submanifolds of $`(\pi ,\mathrm{\Omega })`$. ( Proof ) Let $`S`$ be an integral submanifold of $`𝒟_\alpha `$. First remember that, for every $`𝐩S`$, the $`m`$-orthogonal multisymplectic complement of $`S`$ at $`𝐩`$ is the vector space $$\mathrm{T}_𝐩S^{,m}:=\{X_𝐩\mathrm{T}_𝐩\pi |𝑖(X_𝐩𝒳_𝐩)\mathrm{\Omega }_𝐩=0,\text{for every }𝒳_𝐩^m\mathrm{T}_𝐩S^m(𝒟_\alpha )_𝐩\}$$ and $`S`$ is said to be a $`m`$-coisotropic submanifold of $`(\pi ,\mathrm{\Omega })`$ if $`\mathrm{T}_𝐩S^{,m}\mathrm{T}_𝐩S`$ , . Then, for every $`X_𝐩\mathrm{T}_𝐩S^{,m}`$, if $`𝒳_\alpha `$ is a HDW-multivector field for the extended Hamiltonian system $`(\pi ,\mathrm{\Omega },\alpha )`$, as $`(𝒳_\alpha )_𝐩\mathrm{\Lambda }^m\mathrm{T}_𝐩S`$, by Proposition 5 we have $$0=𝑖(X_𝐩)𝑖((𝒳_\alpha )_𝐩)\mathrm{\Omega }_𝐩=𝑖(X_𝐩)\alpha _𝐩$$ therefore $`X_𝐩(𝒟_\alpha )_𝐩=\mathrm{T}_𝐩S`$. This statement generalizes a well-known result in time-dependent mechanics (see the example in section 6.2): considering the line bundle $`\mu :\mathrm{T}^{}Q\times \mathrm{T}^{}\mathrm{T}^{}Q\times `$, the zero section gives a canonical coisotropic embedding of the submanifold $`\mathrm{T}^{}Q\times `$ into the symplectic manifold $`\mathrm{T}^{}(Q\times )\mathrm{T}^{}Q\times \mathrm{T}^{}`$. Furthermore, in field theories, every maximal integral submanifold $`S`$ of $`𝒟_\alpha `$ gives a local $`m`$-coisotropic embedding of $`U\mu (S)J^1\pi ^{}`$ into $`\pi `$, given by $`(\mu |_S)^1`$, which is obviously not canonical. ### 4.3 Relation between extended and restricted Hamiltonian systems Now we can establish the relation between extended and restricted Hamiltonian systems in $`J^1\pi ^{}`$. Taking into account the considerations made in the above section, we can state: ###### Theorem 5 Let $`(\pi ,\mathrm{\Omega },\alpha )`$ be an extended global Hamiltonian system, and $`(J^1\pi ^{},h)`$ a restricted Hamiltonian system such that $`\mathrm{Im}h=S`$ is an integral submanifold of $`𝒟_\alpha `$. For every $`𝒳_\alpha \text{X}^m(\pi )`$ solution to the equations (8): $$𝑖(𝒳_\alpha )\mathrm{\Omega }=(1)^{m+1}\alpha ,𝑖(𝒳_\alpha )(\overline{\kappa }^{}\omega )=1$$ there exists $`𝒳_h\text{X}^m(J^1\pi ^{})`$ which is $`h`$-related with $`𝒳_\alpha `$ and is a solution to the equations $$𝑖(𝒳_h)\mathrm{\Omega }_h=0,𝑖(𝒳_h)(\overline{\tau }^{}\omega )=1$$ (i.e., satisfying the conditions $`1`$ and $`2`$ in Remark 3). Furthermore, if $`𝒳_\alpha `$ is integrable, then $`𝒳_h`$ is integrable too, and the integral sections of $`𝒳_h`$ are recovered from those of $`𝒳_\alpha `$ as follows: if $`\stackrel{~}{\psi }:M\pi `$ is an integral section of $`𝒳_\alpha `$, then $`\psi =\mu \stackrel{~}{\psi }:MJ^1\pi ^{}`$ is an integral section of $`𝒳_h`$. ( Proof ) Given $`S=\mathrm{Im}h`$, let $`ȷ_S:S\pi `$ be the natural embedding, and $`h_S:J^1\pi ^{}S`$ the diffeomorphism between $`J^1\pi ^{}`$ and $`\mathrm{Im}h`$, then $`h=ȷ_Sh_S`$. If $`𝒳_\alpha \text{X}^m(\pi )`$ is a solution to the equations (8), by Proposition 5, it is tangent to $`S`$, then there exists $`𝒳_S\text{X}^m(S)`$ such that $`\mathrm{\Lambda }^mȷ_S𝒳_S=𝒳_\alpha |_S`$. Let $`𝒳_h\text{X}^m(J^1\pi ^{})`$ defined by $`𝒳_h=\mathrm{\Lambda }^mh_S^1𝒳_S`$. Therefore, from the equation $`𝑖(𝒳_\alpha )\mathrm{\Omega }=(1)^{m+1}\alpha `$ and the condition $`ȷ_S^{}\alpha =0`$ (which holds because $`S`$ is an integral submanifold of $`𝒟_\alpha `$), we obtain $`0`$ $`=`$ $`h^{}[𝑖(𝒳_\alpha )\mathrm{\Omega }(1)^{m+1}\alpha ]=(ȷ_Sh_S)^{}[𝑖(𝒳_\alpha )\mathrm{\Omega }(1)^{m+1}\alpha ]=(h_S^{}ȷ_S^{})[𝑖(𝒳_\alpha )\mathrm{\Omega }(1)^{m+1}\alpha ]`$ $`=`$ $`h_S^{}[𝑖(𝒳_S)ȷ_S^{}\mathrm{\Omega }(1)^{m+1}ȷ_S^{}\alpha ]=𝑖(𝒳_h)(h_S^{}ȷ_S^{})\mathrm{\Omega }=𝑖(𝒳_h)h^{}\mathrm{\Omega }=𝑖(𝒳_h)\mathrm{\Omega }_h`$ Furthermore, bearing in mind that $`\mu h=\mathrm{Id}_{J^1\pi ^{}}`$, we have that $$𝑖(𝒳_h)(\overline{\tau }^{}\omega )=𝑖(𝒳_h)[(\overline{\kappa }h)^{}\omega ]=h^{}[𝑖(𝒳_\alpha )(\overline{\kappa }^{}\omega )]$$ and, if $`𝑖(𝒳_\alpha )(\overline{\kappa }^{}\omega )=1`$, this equality holds, in particular, at the points of the image of $`h`$, therefore $`𝑖(𝒳_h)(\overline{\tau }^{}\omega )=1`$. Then $`𝒳_h`$ is the desired multivector field, since $`𝒳_h|_S=𝒳_S=\mathrm{\Lambda }_{}^m𝒳_h`$. Finally, if $`𝒳_\alpha `$ is integrable, as it is tangent to $`S`$, the integral sections of $`𝒳_\alpha `$ passing through any point of $`S`$ remain in $`S`$, and hence they are the integral sections of $`𝒳_S`$, so $`𝒳_\alpha `$ is integrable and, as a consequence, $`𝒳_h`$ is integrable too. All of these properties lead to establish the following: ###### Definition 6 Given an extended global Hamiltonian system $`(\pi ,\mathrm{\Omega },\alpha )`$, and considering all the Hamiltonian sections $`h:J^1\pi ^{}\pi `$ such that $`\mathrm{Im}h`$ are integral submanifolds of $`𝒟_\alpha `$, we have a family $`\{(J^1\pi ^{},h)\}_\alpha `$, which will be called the class of restricted Hamiltonian systems associated with $`(\pi ,\mathrm{\Omega },\alpha )`$. As it is obvious, in general, the above result holds only locally. The following result show how to obtain extended Hamiltonian systems from restricted Hamiltonian ones, at least locally. In fact: ###### Proposition 7 Given a restricted Hamiltonian system $`(J^1\pi ^{},h)`$, let $`ȷ_S:S=\mathrm{Im}h\pi `$ be the natural embedding. Then, there exists a unique local form $`\alpha \Omega ^1(\pi )`$ such that: 1. $`\alpha Z^1(\pi )`$ (it is a closed form). 2. $`ȷ_S^{}\alpha =0`$. 3. $`𝑖(Y)\alpha 0`$, for every non-vanishing $`Y\text{X}^{\mathrm{V}(\mu )}(\pi )`$ and, in particular, such that $`𝑖\left({\displaystyle \frac{}{p}}\right)\alpha =1`$, for every system of natural coordinates $`(x^\nu ,y^A,p_A^\nu ,p)`$ in $`\pi `$, adapted to the bundle $`\pi :EM`$ (with $`\omega =\mathrm{d}^mx`$). ( Proof ) Suppose that there exist $`\alpha ,\alpha ^{}`$ satisfying the above conditions. Taking into account the comments after Proposition 10, we have that, locally in $`U\pi `$, $`\alpha =\mathrm{d}p+\beta `$ and $`\alpha ^{}=\mathrm{d}p+\beta ^{}`$, where $`\beta =\mu ^{}\overline{\beta }`$, $`\beta ^{}=\mu ^{}\overline{\beta }^{}`$, with $`\overline{\beta },\overline{\beta }^{}B^1(\mu (U))`$ (they are exact $`1`$-forms). From condition 2 in the statement we have that $`ȷ_S^{}\alpha =ȷ_S^{}\alpha ^{}`$; hence $`0=ȷ_S^{}(\alpha \alpha ^{})`$ $``$ $`0=ȷ_S^{}(\beta \beta ^{})=(\mu ȷ_S)^{}(\overline{\beta }\overline{\beta }^{})`$ $``$ $`0=(\mu ȷ_Sh_S)^{}(\overline{\beta }\overline{\beta }^{})\overline{\beta }\overline{\beta }^{}=0\overline{\beta }=\overline{\beta }^{}\overline{\alpha }=\overline{\alpha }^{}`$ since $`\mu ȷ_Sh_S=\mu h=\mathrm{Id}_{\mu (U)}`$. This proves the uniqueness. The existence is trivial since, locally, every section $`h`$ of $`\mu `$ is given by a function $`\mathrm{h}\mathrm{C}^{\mathrm{}}(\mu (U))`$ such that $`p=\mathrm{h}(x^\nu ,y^A,p_A^\nu )`$. Hence $`\alpha |_{\mu (U)}=\mathrm{d}p+\mathrm{d}(\mu ^{}\mathrm{h})\mathrm{d}p+\mathrm{d}\stackrel{~}{\mathrm{h}}`$. ###### Definition 7 Given a restricted Hamiltonian system $`(J^1\pi ^{},h)`$, let $`\alpha \Omega ^1(\pi )`$ be the local form satisfying the conditions in the above proposition. The couple $`(\pi ,\alpha )`$ will be called the (local) extended Hamiltonian system associated with $`(J^1\pi ^{},h)`$. As a consequence of the last proposition, if $`\alpha =\mathrm{d}p+\mu ^{}\overline{\beta }`$, there exists a class $`\{\mathrm{h}\}\mathrm{C}^{\mathrm{}}(\mu (U))/`$, such that $`\overline{\beta }=\mathrm{dh}`$, where $`\mathrm{h}`$ is a representative of this class. Then: ###### Corollary 1 Let $`\alpha `$ be the unique local $`1`$-form verifying the conditions of Proposition 7, associated with a section $`h`$. Consider its characteristic distribution $`𝒟_\alpha `$, and let $`\{h\}_\alpha `$ the family of local sections of $`\mu `$ such that $`\mathrm{Im}hS`$ are local integral submanifolds of $`𝒟_\alpha `$. Then, for every $`h^{}\{h\}_\alpha `$, we have that $`\mathrm{Im}h^{}`$ is locally a level set of the function $`\mathrm{H}=p+\mu ^{}\mathrm{h}p+\stackrel{~}{\mathrm{h}}`$. ( Proof ) If $`S=\mathrm{Im}h^{}`$ then, for every $`\mathrm{p}S`$, we have $`\mathrm{T}_\mathrm{p}S=(𝒟_\alpha )_\mathrm{p}`$, which is equivalent to $`\mathrm{d}(p+\mu ^{}\mathrm{h})|_{\mathrm{T}_\mathrm{p}S}=0`$, and this holds if, and only if, $`\mathrm{H}|_S(p+\mu ^{}\mathrm{h})|_S=ctn.`$ Bearing in mind these considerations, we can finally prove that: ###### Proposition 8 Let $`(\pi ,\mathrm{\Omega },\alpha )`$ be an extended Hamiltonian system, and $`\{(J^1\pi ^{},h)\}_\alpha `$ the class of restricted Hamiltonian systems associated with $`(\pi ,\mathrm{\Omega },\alpha )`$. Consider the submanifolds $`\{S_h=\mathrm{Im}h\}`$, for every Hamiltonian section $`h`$ in this class, and let $`ȷ_{S_h}:S_h\pi `$ be the natural embeddings. Then the submanifolds $`(S_h,ȷ_{S_h}^{}\mathrm{\Omega })`$ are multisymplectomorphic. ( Proof ) Let $`h_1,h_2\{h\}`$ and $`S_1=\mathrm{Im}h_1`$, $`S_2=\mathrm{Im}h_2`$. We have the diagram $$\begin{array}{ccccc}& \text{}& \text{}& \text{}& \\ & \text{}& \text{}& \text{}& \\ \text{}& & \text{}& & \text{}\end{array}$$ Denote $`\mathrm{\Omega }_1=ȷ_{S_1}^{}\mathrm{\Omega }`$, $`\mathrm{\Omega }_2=ȷ_{S_2}^{}\mathrm{\Omega }`$. As a consequence of the above corollary, if $`\mathrm{\Omega }_{h_1}=h_1^{}\mathrm{\Omega }`$, $`\mathrm{\Omega }_{h_2}=h_2^{}\mathrm{\Omega }`$, we have that $`\mathrm{\Omega }_{h_1}=\mathrm{\Omega }_{h_2}`$. But $$\mathrm{\Omega }_{h_1}=\mathrm{\Omega }_{h_2}(ȷ_{S_1}h_{S_1})^{}\mathrm{\Omega }=(ȷ_{S_2}h_{S_2})^{}\mathrm{\Omega }h_{S_1}^{}\mathrm{\Omega }_1=h_{S_2}^{}\mathrm{\Omega }_2$$ Then, the map $`\mathrm{\Phi }:=h_{S_2}\mu _1:S_1S_2`$ is a multisymplectomorphism. In fact, it is obviously a diffeomorphism, and $$\mathrm{\Phi }^{}\mathrm{\Omega }_2=(h_{S_2}\mu _1)^{}\mathrm{\Omega }_2=\mu _1^{}h_{S_2}^{}\mathrm{\Omega }_2=\mu _1^{}h_{S_1}^{}\mathrm{\Omega }_1=(h_{S_1}\mu _1)^{}\mathrm{\Omega }_1=\mathrm{\Omega }_1$$ As an immediate consequence of this, if $`𝒳_\alpha \text{X}^m(\pi )`$ is a solution to the equations (8), the multivector fields $`𝒳_{S_h}\text{X}^m(S_h)`$ such that $`\mathrm{\Lambda }^m(ȷ_{S_h})_{}𝒳_{S_h}=𝒳_\alpha |_S`$, for every submanifold $`S_h`$ of this family (see the proof of Theorem 5), are related by these multisymplectomorphisms. ### 4.4 Variational principle and field equations As in the case of restricted Hamiltonian systems, the field equations for extended Hamiltonian systems can be derived from a suitable variational principle. First, denote by $`\text{X}_\alpha (\pi )`$ the set of vector fields $`Z\text{X}(\pi )`$ which are sections of the subbundle $`𝒟_\alpha `$ of $`\mathrm{T}\pi `$, that is, satisfying that $`𝑖(Z)\alpha =0`$ (and hence, they are tangent to all the integral submanifolds of $`𝒟_\alpha `$). Let $`\text{X}_\alpha ^{\mathrm{V}(\overline{\kappa })}(\pi )\text{X}_\alpha (\pi )`$ be those which are also $`\overline{\kappa }`$-vertical. Furthermore, as we have seen in previous sections, the image of the sections $`\stackrel{~}{\psi }:M\pi `$, which are solutions to the extended field equations, must be in the integral submanifolds of the characteristic distribution $`𝒟_\alpha `$; that is, they are also integral submanifolds, and hence $`ȷ_{\stackrel{~}{\psi }}^{}\alpha =0`$ (where $`ȷ_{\stackrel{~}{\psi }}:\mathrm{Im}\stackrel{~}{\psi }\pi `$ is the natural embedding). We will denote by $`\mathrm{\Gamma }_\alpha (M,\pi )`$ the set of sections of $`\overline{\kappa }`$ satisfying that $`ȷ_{\stackrel{~}{\psi }}^{}\alpha =0`$. Taking all of this into account, we can state the following: ###### Definition 8 Let $`(\pi ,\mathrm{\Omega },\alpha )`$ be an extended Hamiltonian system. Consider the map $$\begin{array}{ccccc}\stackrel{~}{𝐇}_\alpha & :& \mathrm{\Gamma }_\alpha (M,\pi )& & \\ & & \stackrel{~}{\psi }& & _U\stackrel{~}{\psi }^{}\mathrm{\Theta }\end{array}$$ (where the convergence of the integral is assumed). The variational problem for this extended Hamiltonian system is the search for the critical (or stationary) sections of the functional $`\stackrel{~}{𝐇}_\alpha `$, with respect to the variations of $`\stackrel{~}{\psi }\mathrm{\Gamma }_\alpha (M,\pi )`$ given by $`\stackrel{~}{\psi }_t=\sigma _t\stackrel{~}{\psi }`$, where $`\{\sigma _t\}`$ is the local one-parameter group of any compact-supported vector field $`Z\text{X}_\alpha ^{\mathrm{V}(\overline{\kappa })}(\pi )`$, that is $$\frac{d}{dt}|_{t=0}_U\stackrel{~}{\psi }_t^{}\mathrm{\Theta }=0$$ This is the extended Hamilton-Jacobi principle. Observe that, as $`\alpha `$ is closed, the variation of the set $`\mathrm{\Gamma }_\alpha (M,\pi )`$ is stable under the action of $`\text{X}_\alpha ^{\mathrm{V}(\overline{\kappa })}(\pi )`$. In fact; being $`\alpha `$ closed, for every $`Z\text{X}_\alpha ^{\mathrm{V}(\overline{\kappa })}(\pi )`$, we have that $`L(Z)\alpha =𝑖(Z)\mathrm{d}\alpha +\mathrm{d}𝑖(Z)\alpha =0`$, that is, $`\sigma _t^{}\alpha =\alpha `$. Hence, if $`\stackrel{~}{\psi }\mathrm{\Gamma }_\alpha (M,\pi )`$, we obtain $$\stackrel{~}{\psi }_t^{}\alpha =(\sigma _t\stackrel{~}{\psi })^{}\alpha =\stackrel{~}{\psi }^{}\sigma _t^{}\alpha =\stackrel{~}{\psi }^{}\alpha =0$$ Then we have the following fundamental theorems: ###### Theorem 6 Let $`(\pi ,\mathrm{\Omega },\alpha )`$ be an extended Hamiltonian system. The following assertions on a section $`\stackrel{~}{\psi }\mathrm{\Gamma }_\alpha (M,\pi )`$ are equivalent: 1. $`\stackrel{~}{\psi }`$ is a critical section for the variational problem posed by the extended Hamilton-Jacobi principle. 2. $`\stackrel{~}{\psi }^{}𝑖(Z)\mathrm{\Omega }=0`$, for every $`Z\text{X}_\alpha ^{\mathrm{V}(\overline{\kappa })}(\pi )`$. 3. $`\stackrel{~}{\psi }^{}𝑖(X)\mathrm{\Omega }=0`$, for every $`X\text{X}_\alpha (\pi )`$. 4. If $`(U;x^\nu ,y^A,p_A^\nu ,p)`$ is a natural system of coordinates in $`\pi `$, then $`\stackrel{~}{\psi }`$ satisfies the following system of equations in $`U`$ $$\frac{(y^A\stackrel{~}{\psi })}{x^\nu }=\frac{\stackrel{~}{\mathrm{h}}}{p_A^\nu }\stackrel{~}{\psi },\frac{(p_A^\nu \stackrel{~}{\psi })}{x^\nu }=\frac{\stackrel{~}{\mathrm{h}}}{y^A}\stackrel{~}{\psi },\frac{(p\stackrel{~}{\psi })}{x^\nu }=\frac{(\stackrel{~}{\mathrm{h}}\stackrel{~}{\psi })}{x^\nu }$$ (17) where $`\stackrel{~}{\mathrm{h}}=\mu ^{}\mathrm{h}`$, for some $`\mathrm{h}\mathrm{C}^{\mathrm{}}(\mu (U))`$, is any function such that $`\alpha |_U=\mathrm{d}p+\mathrm{d}\stackrel{~}{\mathrm{h}}(x^\nu ,y^A,p_A^\nu )`$. These are the extended Hamilton-De Donder-Weyl equations of the extended Hamiltonian system. ( Proof ) ($`12`$) We assume that $`U`$ is a $`(m1)`$-dimensional manifold and that $`\overline{\kappa }(supp(Z))U`$, for every compact-supported $`Z\text{X}_\alpha ^{\mathrm{V}(\overline{\kappa })}(\pi )`$. Then $`{\displaystyle \frac{d}{dt}}|_{t=0}{\displaystyle _U}\stackrel{~}{\psi }_t^{}\mathrm{\Theta }`$ $`=`$ $`{\displaystyle \frac{d}{dt}}|_{t=0}{\displaystyle _U}\stackrel{~}{\psi }^{}(\sigma _t^{}\mathrm{\Theta })={\displaystyle _U}\stackrel{~}{\psi }^{}\left(\underset{t0}{lim}{\displaystyle \frac{\sigma _t^{}\mathrm{\Theta }\mathrm{\Theta }}{t}}\right)`$ $`=`$ $`{\displaystyle _U}\stackrel{~}{\psi }^{}(L(Z)\mathrm{\Theta })={\displaystyle _U}\stackrel{~}{\psi }^{}(𝑖(Z)\mathrm{d}\mathrm{\Theta }+\mathrm{d}𝑖(Z)\mathrm{\Theta })`$ $`=`$ $`{\displaystyle _U}\stackrel{~}{\psi }^{}(𝑖(Z)\mathrm{\Omega }\mathrm{d}𝑖(Z)\mathrm{\Theta })={\displaystyle _U}\stackrel{~}{\psi }^{}(𝑖(Z)\mathrm{\Omega })+{\displaystyle _U}\mathrm{d}[\stackrel{~}{\psi }^{}(𝑖(Z)\mathrm{\Theta })]`$ $`=`$ $`{\displaystyle _U}\stackrel{~}{\psi }^{}(𝑖(Z)\mathrm{\Omega })+{\displaystyle _U}\stackrel{~}{\psi }^{}(𝑖(Z)\mathrm{\Theta })={\displaystyle _U}\stackrel{~}{\psi }^{}(𝑖(Z)\mathrm{\Omega })`$ (as a consequence of Stoke’s theorem and the hypothesis made on the supports of the vertical fields). Thus, by the fundamental theorem of the variational calculus we conclude that $`{\displaystyle \frac{d}{dt}}|_{t=0}{\displaystyle _U}\psi _t^{}\mathrm{\Theta }=0`$ $``$ $`\stackrel{~}{\psi }^{}(𝑖(Z)\mathrm{\Omega })=0`$, for every compact-supported $`Z\text{X}_\alpha ^{\mathrm{V}(\overline{\kappa })}(\pi )`$. But, as compact-supported vector fields generate locally the $`\mathrm{C}^{\mathrm{}}(\pi )`$-module of vector fields in $`\pi `$, it follows that the last equality holds for every $`Z\text{X}_\alpha ^{\mathrm{V}(\overline{\kappa })}(\pi )`$. ($`23`$) If $`𝐩\mathrm{Im}\stackrel{~}{\psi }`$, and $`S`$ is the integral submanifold of $`𝒟_\alpha `$ passing through $`𝐩`$, then $$(𝒟_\alpha )_𝐩=[\mathrm{V}_𝐩(\overline{\kappa })(𝒟_\alpha )_𝐩]\mathrm{T}_𝐩(\mathrm{Im}\stackrel{~}{\psi })$$ So, for every $`X\text{X}_\alpha (\pi )`$, $$X_𝐩=(X_𝐩\mathrm{T}_𝐩(\stackrel{~}{\psi }\overline{\kappa })(X_𝐩))+\mathrm{T}_𝐩(\stackrel{~}{\psi }\overline{\kappa })(X_𝐩)X_𝐩^V+X_𝐩^{\stackrel{~}{\psi }}$$ and therefore $$\stackrel{~}{\psi }^{}(𝑖(X)\mathrm{\Omega })=\stackrel{~}{\psi }^{}(𝑖(X^V)\mathrm{\Omega })+\stackrel{~}{\psi }^{}(𝑖(X^{\stackrel{~}{\psi }})\mathrm{\Omega })=\stackrel{~}{\psi }^{}(𝑖(X^{\stackrel{~}{\psi }})\mathrm{\Omega })=0$$ since $`\stackrel{~}{\psi }^{}(𝑖(X^V)\mathrm{\Omega })=0`$ by the above item, and furthermore, $`X_𝐩^{\stackrel{~}{\psi }}\mathrm{T}_𝐩(\mathrm{Im}\stackrel{~}{\psi })`$, and $`dim(\mathrm{Im}\stackrel{~}{\psi })=m`$, being $`\mathrm{\Omega }\mathrm{\Omega }^{m+1}(\pi )`$. Hence we conclude that $`\stackrel{~}{\psi }^{}(𝑖(X)\mathrm{\Omega })=0`$, for every $`X\text{X}_\alpha (\pi )`$. The converse is proved reversing this reasoning. ($`34`$) The local expression of any $`X\text{X}_\alpha (\pi )`$ is $$X=\lambda ^\nu \frac{}{x^\nu }+\beta ^A\frac{}{y^A}+\gamma _A^\nu \frac{}{p_A^\nu }\left(\alpha ^\eta \frac{\stackrel{~}{\mathrm{h}}}{x^\eta }+\beta ^B\frac{\stackrel{~}{\mathrm{h}}}{y^B}+\gamma _B^\eta \frac{\stackrel{~}{\mathrm{h}}}{p_B^\eta }\right)\frac{}{p}$$ then, taking into account the local expression (2) of $`\mathrm{\Omega }`$, if $`\stackrel{~}{\psi }=(x^\nu ,y^A(x^\eta ),p_A^\nu (x^\eta ),p(x^\eta ))`$, we obtain $`\stackrel{~}{\psi }^{}𝑖(X)\mathrm{\Omega }`$ $`=`$ $`\lambda ^\eta \left({\displaystyle \frac{(p\stackrel{~}{\psi })}{x^\nu }}+{\displaystyle \frac{\stackrel{~}{\mathrm{h}}}{x^\nu }}|_{\stackrel{~}{\psi }}{\displaystyle \underset{\eta \nu }{}}\left({\displaystyle \frac{\stackrel{~}{\mathrm{h}}}{p_A^\nu }}|_{\stackrel{~}{\psi }}{\displaystyle \frac{(p_A^\eta \stackrel{~}{\psi })}{x^\eta }}{\displaystyle \frac{\stackrel{~}{\mathrm{h}}}{p_A^\eta }}|_{\stackrel{~}{\psi }}{\displaystyle \frac{(p_A^\eta \stackrel{~}{\psi })}{x^\nu }}\right)\right)\mathrm{d}^mx`$ $`+\beta ^A\left({\displaystyle \frac{(p_A^\nu \stackrel{~}{\psi })}{x^\nu }}+{\displaystyle \frac{\stackrel{~}{\mathrm{h}}}{y^A}}|_{\stackrel{~}{\psi }}\right)\mathrm{d}^mx+\gamma _A^\nu \left({\displaystyle \frac{(y^A\stackrel{~}{\psi })}{x^\nu }}+{\displaystyle \frac{\stackrel{~}{\mathrm{h}}}{p_A^\nu }}|_{\stackrel{~}{\psi }}\right)\mathrm{d}^mx`$ and, as this holds for every $`X\text{X}_\alpha (\pi )`$ (i.e., for every $`\lambda ^\eta ,\beta ^A,\gamma _A^\nu `$), we conclude that $`\psi ^{}𝑖(X)\mathrm{\Omega }_h=0`$ if, and only if, $`{\displaystyle \frac{(y^A\stackrel{~}{\psi })}{x^\nu }}`$ $`=`$ $`{\displaystyle \frac{\stackrel{~}{\mathrm{h}}}{p_A^\nu }}|_{\stackrel{~}{\psi }},{\displaystyle \frac{(p_A^\nu \stackrel{~}{\psi })}{x^\nu }}={\displaystyle \frac{\stackrel{~}{\mathrm{h}}}{y^A}}|_{\stackrel{~}{\psi }}`$ (18) $`{\displaystyle \frac{(p\stackrel{~}{\psi })}{x^\nu }}`$ $`=`$ $`{\displaystyle \frac{\stackrel{~}{\mathrm{h}}}{x^\nu }}|_{\stackrel{~}{\psi }}+{\displaystyle \underset{\eta \nu }{}}\left({\displaystyle \frac{\stackrel{~}{\mathrm{h}}}{p_A^\nu }}|_{\stackrel{~}{\psi }}{\displaystyle \frac{(p_A^\eta \stackrel{~}{\psi })}{x^\eta }}{\displaystyle \frac{\stackrel{~}{\mathrm{h}}}{p_A^\eta }}|_{\stackrel{~}{\psi }}{\displaystyle \frac{(p_A^\eta \stackrel{~}{\psi })}{x^\nu }}\right)`$ (19) and using equations (18) in (19) we obtain $`{\displaystyle \frac{(p\stackrel{~}{\psi })}{x^\nu }}`$ $`=`$ $`{\displaystyle \frac{\stackrel{~}{\mathrm{h}}}{x^\nu }}|_{\stackrel{~}{\psi }}+{\displaystyle \frac{\stackrel{~}{\mathrm{h}}}{p_A^\nu }}|_{\stackrel{~}{\psi }}\left({\displaystyle \frac{(p_A^\delta \stackrel{~}{\psi })}{x^\delta }}{\displaystyle \frac{(p_A^\nu \stackrel{~}{\psi })}{x^\nu }}\right)\left({\displaystyle \frac{\stackrel{~}{\mathrm{h}}}{p_A^\delta }}|_{\stackrel{~}{\psi }}{\displaystyle \frac{(p_A^\delta \stackrel{~}{\psi })}{x^\nu }}{\displaystyle \frac{\stackrel{~}{\mathrm{h}}}{p_A^\nu }}|_{\stackrel{~}{\psi }}{\displaystyle \frac{(p_A^\nu \stackrel{~}{\psi })}{x^\nu }}\right)`$ $`=`$ $`{\displaystyle \frac{\stackrel{~}{\mathrm{h}}}{x^\nu }}|_{\stackrel{~}{\psi }}{\displaystyle \frac{\stackrel{~}{\mathrm{h}}}{y^A}}|_{\stackrel{~}{\psi }}{\displaystyle \frac{(y^A\stackrel{~}{\psi })}{x^\nu }}{\displaystyle \frac{\stackrel{~}{\mathrm{h}}}{p_A^\delta }}|_{\stackrel{~}{\psi }}{\displaystyle \frac{(p_A^\delta \stackrel{~}{\psi })}{x^\nu }}`$ $`=`$ $`{\displaystyle \frac{(\stackrel{~}{\mathrm{h}}\stackrel{~}{\psi })}{x^\nu }}(\delta =1,\mathrm{},m;\nu \mathrm{fixed})`$ ###### Remark 10 It is important to point out that the last group of equations (17) are consistency conditions with respect to the hypothesis made on the sections $`\stackrel{~}{\psi }`$. In fact, this group of equations leads to $`p\stackrel{~}{\psi }=\stackrel{~}{\mathrm{h}}\stackrel{~}{\psi }+ctn.`$ or, what means the same thing, $`\stackrel{~}{\psi }\mathrm{\Gamma }_\alpha (M,\pi )`$. (See also the comment in Remark 9). The rest of equations (17) are just the Hamilton-De Donder-Weyl equations (5) of the restricted case, since the local expressions of the functions $`\stackrel{~}{\mathrm{h}}`$ and $`\mathrm{h}`$ are the same. ###### Theorem 7 Let $`(\pi ,\mathrm{\Omega },\alpha )`$ be an extended Hamiltonian system, and $`𝒳\text{X}^m(\pi )`$ an integrable multivector field verifying the condition $`𝑖(𝒳)(\overline{\kappa }^{}\omega )=1`$. Then, the integral manifolds of $`𝒳`$ are critical section for the variational problem posed by the extended Hamilton-Jacobi principle if, and only if, $`𝒳`$ satisfies the condition $`𝑖(𝒳)\mathrm{\Omega }=(1)^{m+1}\alpha `$. ( Proof ) ($``$) Let $`S`$ be an integral submanifold of $`𝒳`$. By the $`\overline{\kappa }`$-transversality condition $`𝑖(𝒳)(\overline{\kappa }^{}\omega )=1`$, $`S`$ is locally a section of $`\overline{\kappa }`$. Then, for every $`𝐩S`$, there are an open set $`UM`$, with $`\overline{\kappa }(𝐩)U`$, and a local section $`\stackrel{~}{\psi }:UM\pi `$ of $`\overline{\kappa }`$, such that $`\mathrm{Im}\stackrel{~}{\psi }=S|_{\overline{\kappa }^1(U)}`$. Now, let $`qU`$, and $`u_1,\mathrm{},u_m\mathrm{T}_qM`$, with $`𝑖(u_1\mathrm{}u_m)(\omega (\overline{\kappa }(𝐩)))=1`$. Then, there exists $`\lambda `$ such that $$\stackrel{~}{\psi }_{}(u_1\mathrm{}u_m)=\lambda 𝒳_{(\stackrel{~}{\psi }(q))}$$ therefore $$𝑖(\stackrel{~}{\psi }_{}(u_1\mathrm{}u_m))(\mathrm{\Omega }(\stackrel{~}{\psi }(q)))=\lambda (1)^{m+1}\alpha (\stackrel{~}{\psi }(q))$$ Thus, for every $`X\text{X}_\alpha (\pi )`$ we obtain that $$𝑖(X_{(\stackrel{~}{\psi }(q))})𝑖(\stackrel{~}{\psi }_{}(u_1\mathrm{}u_m)(\mathrm{\Omega }(\stackrel{~}{\psi }(q)))=0$$ hence $`\stackrel{~}{\psi }^{}𝑖(X)\mathrm{\Omega }=0`$, for every $`X\text{X}_\alpha (\pi )`$, and $`\stackrel{~}{\psi }`$ is a critical section by the third item of the last Theorem. ($``$) Let $`𝐩\pi `$, by the hypothesis there exists a section $`\stackrel{~}{\psi }:M\pi `$ such that 1. $`\stackrel{~}{\psi }(𝐩)=\overline{\kappa }(𝐩)`$. 2. $`\stackrel{~}{\psi }`$ is a critical section for the extended Hamilton-Jacobi variational problem, that is, $`\stackrel{~}{\psi }^{}𝑖(X)\mathrm{\Omega }=0`$, for every $`X\text{X}_\alpha (\pi )`$. 3. $`\mathrm{Im}\stackrel{~}{\psi }`$ is an integral submanifold of $`𝒳`$. Now, let $`u_1,\mathrm{},u_m\mathrm{T}_{\overline{\kappa }(𝐩)}M`$, with $`𝑖(u_1\mathrm{}u_m)\omega (\overline{\kappa }(𝐩)))=1`$. Then, there exists $`\lambda `$ such that $$\stackrel{~}{\psi }_{}(u_1\mathrm{}u_m)=\lambda 𝒳_𝐩$$ but the condition imposed to $`u_1,\mathrm{},u_m`$ leads to $`\lambda =1`$. Therefore $$𝑖(\stackrel{~}{\psi }_{}(u_1\mathrm{}u_m))(\mathrm{\Omega }(𝐩))=𝑖(𝒳_𝐩)(\mathrm{\Omega }(𝐩))$$ Thus, for every $`X\text{X}_\alpha (\pi )`$, as $`\stackrel{~}{\psi }`$ is a critical section, we obtain that $$𝑖(X_𝐩)𝑖(𝒳_𝐩)(\mathrm{\Omega }(𝐩))=0$$ and hence $`𝑖(X)𝑖(𝒳)\mathrm{\Omega }=0`$, for every $`X\text{X}_\alpha (\pi )`$. This implies that $`𝑖(𝒳)\mathrm{\Omega }=f\alpha `$, for some non-vanishing $`f\mathrm{C}^{\mathrm{}}(\pi )`$. Nevertheless, as $`(\pi ,\mathrm{\Omega },\alpha )`$ is an extended Hamiltonian system, in any local chart we have that $`\alpha =\mathrm{d}p+\mathrm{d}\stackrel{~}{\mathrm{h}}(x^\nu ,y^A,p_A^\nu )`$, which by the condition $`𝑖(𝒳)(\overline{\kappa }^{}\omega )=1`$, and bearing in mind the local expression of $`\mathrm{\Omega }`$, leads to $`f=(1)^{m+1}`$. So the result holds. Observe that the extended Hamilton-De Donder-Weyl equations (17) can also be obtained as a consequence of this last theorem, taking into account equations (13), (14), (15) and (16). ## 5 Almost-regular Hamiltonian systems There are many interesting cases in Hamiltonian field theories where the Hamiltonian field equations are established not in $`J^1\pi ^{}`$, but rather in a submanifold of $`J^1\pi ^{}`$ (for instance, when considering the Hamiltonian formalism associated with a singular Lagrangian). Next we consider this kind of systems in $`J^1\pi ^{}`$, as well as in $`\pi `$. ### 5.1 Restricted almost-regular Hamiltonian systems ###### Definition 9 A restricted almost-regular Hamiltonian system is a triple $`(J^1\pi ^{},𝒫,h_𝒫)`$, where: 1. $`𝒫`$ is a submanifold of $`J^1\pi ^{}`$ with $`dim𝒫>n+m`$, and such that, if $`ȷ_𝒫:𝒫J^1\pi ^{}`$ denotes the natural embedding, the map $`\tau _𝒫=\tau ȷ_𝒫:𝒫E`$ is a surjective submersion (and hence, so is the map $`\overline{\tau }_𝒫=\overline{\tau }ȷ_𝒫=\pi \tau _𝒫:𝒫M`$). 2. $`h_𝒫:𝒫\pi `$ satisfies that $`\mu h_𝒫=ȷ_𝒫`$, and it is called a Hamiltonian section of $`\mu `$ on $`𝒫`$. Then, the differentiable forms $$\mathrm{\Theta }_{h_𝒫}:=h_𝒫^{}\mathrm{\Theta },\mathrm{\Omega }_{h_𝒫}:=\mathrm{d}\mathrm{\Theta }_{h_𝒫}=h_𝒫^{}\mathrm{\Omega }$$ are the Hamilton-Cartan $`m`$ and $`(m+1)`$ forms on $`𝒫`$ associated with the Hamiltonian section $`h_𝒫`$. We have the diagram $$\begin{array}{ccc}& \text{}& \\ 𝒫& \text{}& J^1\pi ^{}\\ & \text{}& \end{array}$$ ###### Remark 11 Notice that $`\mathrm{\Omega }_{h_𝒫}`$ is, in general, a $`1`$-degenerate form and hence it is premultisymplectic. This is the main difference with the regular case. Furthermore, if we make the additional assumption that $`𝒫E`$ is a fiber bundle, the Hamilton-Jacobi variational principle of Definition 2 can be stated in the same way, now using sections of $`\overline{\tau }_𝒫:𝒫M`$, and the form $`\mathrm{\Theta }_{h_𝒫}`$. So we look for sections $`\psi _𝒫\mathrm{\Gamma }(M,𝒫)`$ which are stationary with respect to the variations given by $`\psi _t=\sigma _t\psi _𝒫`$, where $`\{\sigma _t\}`$ is a local one-parameter group of any compact-supported $`\overline{\tau }_𝒫`$-vertical vector field $`Z_𝒫\text{X}(𝒫)`$; i.e., such that $$\frac{d}{dt}|_{t=0}_M\psi _t^{}\mathrm{\Theta }_{h_𝒫}=0$$ Then these critical sections will be characterized by the condition (analogous to Theorem 1) $$\psi _𝒫^{}𝑖(X_𝒫)\mathrm{\Omega }_{h_𝒫}=0;\text{for every }X_𝒫\text{X}(𝒫)$$ And, as in the case of restricted Hamiltonian systems (Theorem 7), we have that: ###### Theorem 8 The critical sections of the Hamilton-Jacobi principle are the integral sections $`\psi _𝒫\mathrm{\Gamma }(M,𝒫)`$ of a class of integrable and $`\overline{\tau }_𝒫`$-transverse multivector fields $`\{𝒳_{h_𝒫}\}\text{X}^m(𝒫)`$ satisfying that $$𝑖(𝒳_{h_𝒫})\mathrm{\Omega }_{h_𝒫}=0,\text{for every }𝒳_{h_𝒫}\{𝒳_{h_𝒫}\}$$ or equivalently, the integral sections of an integrable multivector field $`𝒳_{h_𝒫}\text{X}^m(𝒫)`$ such that: 1. $`𝑖(𝒳_{h_𝒫})\mathrm{\Omega }_{h_𝒫}=0`$. 2. $`𝑖(𝒳_{h_𝒫})(\overline{\tau }_𝒫^{}\omega )=1`$. A multivector field $`𝒳_{h_𝒫}\text{X}^m(𝒫)`$ will be called a Hamilton-De Donder-Weyl multivector field for the system $`(J^1\pi ^{},𝒫,h_𝒫)`$ if it is $`\overline{\tau }_𝒫`$-transverse, locally decomposable and verifies the equation $`𝑖(𝒳_{h_𝒫})\mathrm{\Omega }_{h_𝒫}=0`$. Then, the associated connection $`_{h_𝒫}`$, which is a connection along the submanifold $`𝒫`$ (see , and ), is called a Hamilton-De Donder-Weyl connection for $`(J^1\pi ^{},𝒫,h_𝒫)`$, and satisfies the equation $$𝑖(_{h_𝒫})\mathrm{\Omega }_{h_𝒫}=(m1)\mathrm{\Omega }_{h_𝒫}$$ ###### Remark 12 It should be noted that, as $`\mathrm{\Omega }_{h_𝒫}`$ can be $`1`$-degenerate, the existence of the corresponding Hamilton-De Donder-Weyl multivector fields for $`(J^1\pi ^{},𝒫,h_𝒫)`$ is in general not assured except perhaps on some submanifold $`S`$ of $`𝒫`$, where the solution is not unique. A geometric algorithm for determining this submanifold $`S`$ has been developed . ### 5.2 Extended almost-regular Hamiltonian systems ###### Definition 10 An extended almost-regular Hamiltonian system is a triple $`(\pi ,\stackrel{~}{𝒫},\alpha _{\stackrel{~}{𝒫}})`$, such that: 1. $`\stackrel{~}{𝒫}`$ is a submanifold of $`\pi `$ and, if $`ȷ_{\stackrel{~}{𝒫}}:\stackrel{~}{𝒫}\pi `$ denotes the natural embedding, then: 1. $`\kappa _{\stackrel{~}{𝒫}}=\kappa ȷ_{\stackrel{~}{𝒫}}:\stackrel{~}{𝒫}E`$ is a surjective submersion (and hence, so is the map $`\overline{\kappa }_{\stackrel{~}{𝒫}}=\overline{\kappa }ȷ_{\stackrel{~}{𝒫}}=\pi \kappa _{\stackrel{~}{𝒫}}:\stackrel{~}{𝒫}M`$). 2. $`(\mu ȷ_{\stackrel{~}{𝒫}})(\stackrel{~}{𝒫})𝒫`$ is a submanifold of $`J^1\pi ^{}`$. 3. $`\stackrel{~}{𝒫}=\pi |_{\mu (\stackrel{~}{𝒫})}`$; that is, for every $`𝐩\stackrel{~}{𝒫}`$ we have that $`\mu ^1(\mu (𝐩))=𝐩+\mathrm{\Lambda }_1^m\mathrm{T}_{\kappa (𝐩)}^{}E\stackrel{~}{𝒫}`$. 2. $`\alpha _{\stackrel{~}{𝒫}}Z^1(\stackrel{~}{𝒫})`$ (it is a closed $`1`$-form in $`\stackrel{~}{𝒫}`$). 3. There exists a locally decomposable multivector field $`𝒳_{\alpha _{\stackrel{~}{𝒫}}}\text{X}^m(\stackrel{~}{𝒫})`$ satisfying that $$𝑖(𝒳_{\alpha _{\stackrel{~}{𝒫}}})\mathrm{\Omega }_{\stackrel{~}{𝒫}}=(1)^{m+1}\alpha _{\stackrel{~}{𝒫}},𝑖(𝒳_{\alpha _{\stackrel{~}{𝒫}}})(\overline{\kappa }_{\stackrel{~}{𝒫}}^{}\omega )=1\text{(}\overline{\kappa }_{\stackrel{~}{𝒫}}\text{-transversality)}$$ (20) where $`\mathrm{\Omega }_{\stackrel{~}{𝒫}}=ȷ_{\stackrel{~}{𝒫}}^{}\mathrm{\Omega }`$. If $`\alpha _{\stackrel{~}{𝒫}}`$ is an exact form, then $`(\pi ,\stackrel{~}{𝒫},\alpha _{\stackrel{~}{𝒫}})`$ is an extended almost-regular global Hamiltonian system. In this case there exist functions $`\mathrm{H}_{\stackrel{~}{𝒫}}\mathrm{C}^{\mathrm{}}(\stackrel{~}{𝒫})`$, which are called Hamiltonian functions of the system, such that $`\alpha _{\stackrel{~}{𝒫}}=\mathrm{dH}_{\stackrel{~}{𝒫}}`$. (For an extended Hamiltonian system, these functions exist only locally, and they are called local Hamiltonian functions). ###### Remark 13 As straighforward consequences of this definition we have that: * The condition (1.c) of Definition 10 imply, in particular, that $`dim\stackrel{~}{𝒫}>dimE+1`$. Furthermore, it means that $`\stackrel{~}{𝒫}`$ is the union of fibers of $`\mu `$. * $`\kappa ȷ_{\stackrel{~}{𝒫}}`$ is a surjective submersion if, and only if, so is $`\tau \mu ȷ_{\stackrel{~}{𝒫}}`$. This means that $`𝒫\mathrm{Im}(\mu ȷ_{\stackrel{~}{𝒫}})`$ is a submanifold verifying the conditions stated in the first item of definition 9, and such that $`dim𝒫=dim\stackrel{~}{𝒫}1`$, as a consequence of the properties given in item 1 of Definition 10. This submanifold is diffeomorphic to $`\stackrel{~}{𝒫}/\mathrm{\Lambda }_1^m\mathrm{T}^{}E`$. Denoting $`\mu _{\stackrel{~}{𝒫}}=\mu ȷ_{\stackrel{~}{𝒫}}:\stackrel{~}{𝒫}J^1\pi ^{}`$, and $`\stackrel{~}{\mu }_{\stackrel{~}{𝒫}}:\stackrel{~}{𝒫}𝒫`$ its restriction to the image (that is, such that $`\mu _{\stackrel{~}{𝒫}}=ȷ_𝒫\stackrel{~}{\mu }_{\stackrel{~}{𝒫}}`$), we have the diagram $$\begin{array}{ccc}𝒫& \text{}& J^1\pi ^{}\\ \text{}& \text{}& \text{}\\ \stackrel{~}{𝒫}& \text{}& \pi \\ & \text{}& \end{array}$$ ###### Remark 14 In addition, as for extended Hamiltonian systems (see Remarks 5 and 6), the integrability of $`𝒳_{\alpha _{\stackrel{~}{𝒫}}}`$ is not assured, so it must be imposed. Then all the multivector fields in the integrable class $`\{𝒳_{\alpha _{\stackrel{~}{𝒫}}}\}`$ have the same integral sections. As in Propositions 10 and 2, we have that: ###### Proposition 9 If $`(\pi ,\stackrel{~}{𝒫},\alpha _{\stackrel{~}{𝒫}})`$ is an extended almost-regular Hamiltonian system, then $`𝑖(Y_{\stackrel{~}{𝒫}})\alpha _{\stackrel{~}{𝒫}}0`$, for every non-vanishing $`\mu _{\stackrel{~}{𝒫}}`$-vertical vector field $`Y_{\stackrel{~}{𝒫}}\text{X}^{\mathrm{V}(\mu _{\stackrel{~}{𝒫}})}(\stackrel{~}{𝒫})`$. In particular, for every system of natural coordinates in $`\stackrel{~}{𝒫}`$ adapted to the bundle $`\pi :EM`$ (with $`\omega =\mathrm{d}^mx`$), $$𝑖\left(\frac{}{p}\right)\alpha _{\stackrel{~}{𝒫}}=1$$ ( Proof ) As a consequence of the condition (1.c) of Definition 10, we have the local expression $$\mathrm{\Omega }_{\stackrel{~}{𝒫}}=ȷ_{\stackrel{~}{𝒫}}^{}\mathrm{\Omega }=ȷ_{\stackrel{~}{𝒫}}^{}(\mathrm{d}p_A^\nu \mathrm{d}y^A\mathrm{d}^{m1}x_\nu \mathrm{d}p\mathrm{d}^mx)=ȷ_{\stackrel{~}{𝒫}}^{}(\mathrm{d}p_A^\nu )\mathrm{d}y^A\mathrm{d}^{m1}x_\nu \mathrm{d}p\mathrm{d}^mx$$ and $`Y_{\stackrel{~}{𝒫}}=f{\displaystyle \frac{}{p}}`$, for every $`\mu `$-vertical vector field in $`\stackrel{~}{𝒫}`$. Therefore, the proof follows the same pattern as in the proof of Proposition 10. The last part of the proof is a consequence of the condition (1.c) given in Definition 10, from which we have that every system of natural coordinates in $`\stackrel{~}{𝒫}`$ adapted to the bundle $`\pi :EM`$ contains the coordinate $`p`$ of the fibers of $`\mu `$, and the coordinates $`(x^\nu )`$ in $`E`$. This happens because $`\stackrel{~}{𝒫}`$ reduces only degrees of freedom in the coordinates $`p_A^\nu `$ of $`\pi `$. ###### Proposition 10 If $`(\pi ,\stackrel{~}{𝒫},\alpha _{\stackrel{~}{𝒫}})`$ is an extended Hamiltonian system, locally $`\alpha _{\stackrel{~}{𝒫}}=\mathrm{d}p+\beta _{\stackrel{~}{𝒫}}`$, where $`\beta _{\stackrel{~}{𝒫}}`$ is a closed and $`\stackrel{~}{\mu }_{\stackrel{~}{𝒫}}`$-basic local $`1`$-form in $`\stackrel{~}{𝒫}`$. A multivector field $`𝒳_{\alpha _{\stackrel{~}{𝒫}}}\text{X}^m(\stackrel{~}{𝒫})`$ will be called an extended Hamilton-De Donder-Weyl multivector field for the system $`(\pi ,\stackrel{~}{𝒫},\alpha _{\stackrel{~}{𝒫}})`$ if it is $`\overline{\kappa }_{\stackrel{~}{𝒫}}`$-transverse, locally decomposable and verifies the equation $`𝑖(𝒳_{\alpha _{\stackrel{~}{𝒫}}})\mathrm{\Omega }_{\stackrel{~}{𝒫}}=(1)^{m+1}\alpha _{\stackrel{~}{𝒫}}`$. Then, the associated connection $`_{\alpha _{\stackrel{~}{𝒫}}}`$ is called a Hamilton-De Donder-Weyl connection for $`(\pi ,\stackrel{~}{𝒫},\alpha _{\stackrel{~}{𝒫}})`$, ###### Remark 15 Notice that $`\mathrm{\Omega }_{\stackrel{~}{𝒫}}`$ is usually a $`1`$-degenerate form and hence premultisymplectic. As a consequence, the existence of extended Hamilton-De Donder-Weyl multivector fields for $`(\pi ,\stackrel{~}{𝒫},\alpha _{\stackrel{~}{𝒫}})`$ is not assured, except perhaps on some submanifold $`\stackrel{~}{S}`$ of $`𝒫`$, where the solution is not unique. ### 5.3 Geometric properties of extended almost-regular Hamiltonian systems. Variational principle Let $`(\pi ,\stackrel{~}{𝒫},\alpha _{\stackrel{~}{𝒫}})`$ be an extended almost-regular Hamiltonian system, and the submanifold $`\mu _{\stackrel{~}{𝒫}}(\stackrel{~}{𝒫})𝒫`$. As for the general case, we can define the characteristic distribution $`𝒟_{\alpha _{\stackrel{~}{𝒫}}}`$ of $`\alpha _{\stackrel{~}{𝒫}}`$. Then, following the same pattern as in the proofs of the propositions and theorems given in Section 4.2 we can prove that: ###### Proposition 11 1. $`𝒟_{\alpha _{\stackrel{~}{𝒫}}}`$ is an involutive and $`\mu _{\stackrel{~}{𝒫}}`$-transverse distribution of corank equal to $`1`$. 2. The integral submanifolds of $`𝒟_{\alpha _{\stackrel{~}{𝒫}}}`$ are $`\mu _{\stackrel{~}{𝒫}}`$-transverse submanifolds of $`\pi `$, with dimension equal to $`dim\stackrel{~}{𝒫}1`$. (We denote by $`\stackrel{~}{ȷ}_S:S\stackrel{~}{𝒫}`$ the natural embedding). 3. For every $`𝐩\stackrel{~}{𝒫}`$, we have that $`\mathrm{T}_𝐩\stackrel{~}{𝒫}=\mathrm{V}_𝐩(\mu _{\stackrel{~}{𝒫}})(𝒟_{\alpha _{\stackrel{~}{𝒫}}})_𝐩`$, and thus, in this way, $`\alpha _{\stackrel{~}{𝒫}}`$ defines a connection in the bundle $`\mu _{\stackrel{~}{𝒫}}:\stackrel{~}{𝒫}𝒫`$. 4. If $`S`$ is an integral submanifold of $`𝒟_{\alpha _{\stackrel{~}{𝒫}}}`$, then $`\stackrel{~}{\mu }_{\stackrel{~}{𝒫}}|_S:S𝒫`$ is a local diffeomorphism. 5. For every integral submanifold $`S`$ of $`𝒟_{\alpha _{\stackrel{~}{𝒫}}}`$, and $`𝐩S`$, there exists $`W\stackrel{~}{𝒫}`$, with $`𝐩W`$, such that $`h=(\mu |_{WS})^1`$ is a local Hamiltonian section of $`\stackrel{~}{\mu }_{\stackrel{~}{𝒫}}`$ defined on $`\stackrel{~}{\mu }_{\stackrel{~}{𝒫}}(WS)`$. If $`(\pi ,\stackrel{~}{𝒫},\alpha _{\stackrel{~}{𝒫}})`$ is an extended almost-regular Hamiltonian system, as $`\alpha _{\stackrel{~}{𝒫}}=\mathrm{dH}_{\stackrel{~}{𝒫}}`$ (locally), every local Hamiltonian function $`\mathrm{H}_{\stackrel{~}{𝒫}}`$ is a constraint defining the local integral submanifolds of $`𝒟_{\stackrel{~}{𝒫}}`$. If $`(\pi ,\stackrel{~}{𝒫},\alpha _{\stackrel{~}{𝒫}})`$ is an extended almost-regular global Hamiltonian system, the Hamiltonian functions $`\mathrm{H}_{\stackrel{~}{𝒫}}`$ are globally defined, and we have: ###### Proposition 12 Let $`(\pi ,\stackrel{~}{𝒫},\alpha _{\stackrel{~}{𝒫}})`$ be an extended almost-regular global Hamiltonian system. If there is a global Hamiltonian function $`\mathrm{H}_{\stackrel{~}{𝒫}}\mathrm{C}^{\mathrm{}}(\stackrel{~}{𝒫})`$, and $`k`$, such that $`\mu (\mathrm{H}_{\stackrel{~}{𝒫}}^{}{}_{}{}^{1}(k))=𝒫`$, then there exists a global Hamiltonian section $`h_{\stackrel{~}{𝒫}}\mathrm{\Gamma }(𝒫,\stackrel{~}{𝒫})`$. ###### Proposition 13 Given an extended almost-regular Hamiltonian system $`(\pi ,\stackrel{~}{𝒫},\alpha _{\stackrel{~}{𝒫}})`$ , every extended HDW multivector field $`𝒳_{\alpha _{\stackrel{~}{𝒫}}}\text{X}^m(\stackrel{~}{𝒫})`$ for the system $`(\pi ,\stackrel{~}{𝒫},\alpha _{\stackrel{~}{𝒫}})`$ is tangent to every integral submanifold of $`𝒟_{\alpha _{\stackrel{~}{𝒫}}}`$. At this point, the extended Hamilton-Jacobi variational principle of Definition 8 can be stated in the same way, now using sections $`\psi _{\stackrel{~}{𝒫}}`$ of $`\overline{\kappa }_{\stackrel{~}{𝒫}}:\stackrel{~}{𝒫}M`$, satisfying that $`ȷ_{\psi _{\stackrel{~}{𝒫}}}^{}\alpha _{\stackrel{~}{𝒫}}=0`$ (where $`ȷ_{\psi _{\stackrel{~}{𝒫}}}:\mathrm{Im}\psi _{\stackrel{~}{𝒫}}\stackrel{~}{𝒫}`$ denotes the natural embedding). Thus, using the notation introduced in section 4.4, we look for sections $`\psi _{\stackrel{~}{𝒫}}\mathrm{\Gamma }_{\alpha _{\stackrel{~}{𝒫}}}(M,\stackrel{~}{𝒫})`$ which are stationary with respect to the variations given by $`\stackrel{~}{\psi }_t=\sigma _t\psi _{\stackrel{~}{𝒫}}`$, where $`\{\sigma _t\}`$ is a local one-parameter group of every compact-supported $`Z\text{X}_{\alpha _{\stackrel{~}{𝒫}}}^{\mathrm{V}(\overline{\kappa }_{\stackrel{~}{𝒫}})}(\stackrel{~}{𝒫})`$; that is $$\frac{d}{dt}|_{t=0}_U\stackrel{~}{\psi }_t^{}\mathrm{\Theta }_{\stackrel{~}{𝒫}}=0$$ And then the statements analogous to Theorems 6 and 7 can be established and proven in the present case. ### 5.4 Relation between extended and restricted almost-regular Hamiltonian systems Finally, we study the relation between extended and restricted almost-regular Hamiltonian systems. (The proofs of the following propositions and theorems are analogous to those in Section 4.3). First, bearing in mind Remark 13, we have: ###### Theorem 9 Let $`(\pi ,\stackrel{~}{𝒫},\alpha _{\stackrel{~}{𝒫}})`$ be an extended global Hamiltonian system, and $`(J^1\pi ^{},𝒫,h_𝒫)`$ a restricted Hamiltonian system such that $`dim\stackrel{~}{𝒫}=dim𝒫+1`$, and $`\mathrm{Im}h_𝒫=S`$ is an integral submanifold of $`𝒟_{\alpha _{\stackrel{~}{𝒫}}}`$. Then, for every $`𝒳_{\alpha _{\stackrel{~}{𝒫}}}\text{X}^m(\stackrel{~}{𝒫})`$ solution to the equations: $$𝑖(𝒳_{\alpha _{\stackrel{~}{𝒫}}})\mathrm{\Omega }_{\stackrel{~}{𝒫}}=(1)^{m+1}\alpha _{\stackrel{~}{𝒫}},𝑖(𝒳_{\alpha _{\stackrel{~}{𝒫}}})(\overline{\kappa }_{\stackrel{~}{𝒫}}^{}\omega )=1$$ (i.e., an extended HDW multivector field for $`(\pi ,\stackrel{~}{𝒫},\alpha _{\stackrel{~}{𝒫}})`$) there exists $`𝒳_{h_𝒫}\text{X}^m(𝒫)`$ which is $`h_𝒫`$-related with $`𝒳_\alpha `$ and is a solution to the equations $$𝑖(𝒳_{h_𝒫})\mathrm{\Omega }_{h_𝒫}=0,𝑖(𝒳_{h_𝒫})(\overline{\tau }_𝒫^{}\omega )=1$$ (i.e., a HDW multivector field for $`(J^1\pi ^{},𝒫,h_𝒫)`$). Furthermore, if $`𝒳_{\alpha _{\stackrel{~}{𝒫}}}`$ is integrable, then $`𝒳_{h_𝒫}`$ is integrable too. As a consequence, the following definition can be established: ###### Definition 11 Given an extended almost-regular global Hamiltonian system $`(\pi ,\stackrel{~}{𝒫},\alpha _{\stackrel{~}{𝒫}})`$, and considering all the Hamiltonian sections $`h_𝒫:𝒫\pi `$ such that $`\mathrm{Im}h_𝒫`$ are integral submanifolds of $`𝒟_{\alpha _𝒫}`$, we have a family $`\{(J^1\pi ^{},𝒫,h_𝒫)\}_{\alpha _{\stackrel{~}{𝒫}}}`$, which will be called the class of restricted almost-regular Hamiltonian systems associated with $`(\pi ,\stackrel{~}{𝒫},\alpha _{\stackrel{~}{𝒫}})`$. ###### Remark 16 Observe that, for every Hamiltonian section $`h_𝒫`$ in this class, $`\mathrm{Im}h_𝒫`$ is a submanifold of $`\stackrel{~}{𝒫}`$. Therefore we have induced a Hamiltonian section $`h_{\stackrel{~}{𝒫}}:𝒫\stackrel{~}{𝒫}`$ of $`\stackrel{~}{\mu }_{\stackrel{~}{𝒫}}`$ such that $`h_𝒫=ȷ_{\stackrel{~}{𝒫}}h_{\stackrel{~}{𝒫}}`$. ###### Proposition 14 Let $`(J^1\pi ^{},𝒫,h_𝒫)`$ be a restricted almost-regular Hamiltonian system. 1. There exits a unique submanifold $`\stackrel{~}{𝒫}`$ of $`\pi `$ satisfying the conditions of Definition 10, and such that $`\mu _{\stackrel{~}{𝒫}}(\stackrel{~}{𝒫})=𝒫`$. 2. Consider the submanifold $`S=\mathrm{Im}h_𝒫`$ and the natural embedding $`\stackrel{~}{ȷ}_S:S=\mathrm{Im}h_𝒫\stackrel{~}{𝒫}`$. Then, there exists a unique local form $`\alpha _{\stackrel{~}{𝒫}}\mathrm{\Omega }^1(\stackrel{~}{𝒫})`$ such that: 1. $`\alpha _{\stackrel{~}{𝒫}}Z^1(\stackrel{~}{𝒫})`$ (it is a closed form). 2. $`\stackrel{~}{ȷ}_S^{}\alpha _{\stackrel{~}{𝒫}}=0`$. 3. $`𝑖(Y_{\stackrel{~}{𝒫}})\alpha _{\stackrel{~}{𝒫}}0`$, for every non-vanishing $`Y_{\stackrel{~}{𝒫}}\text{X}^{\mathrm{V}(\mu _{\stackrel{~}{𝒫}})}(\stackrel{~}{𝒫})`$ and, in particular, such that $`𝑖\left({\displaystyle \frac{}{p}}\right)\alpha _{\stackrel{~}{𝒫}}=1`$, for every system of natural coordinates in $`\stackrel{~}{𝒫}`$, adapted to the bundle $`\pi :EM`$ (with $`\omega =\mathrm{d}^mx`$). ( Proof ) The existence and uniqueness of the submanifold $`\stackrel{~}{𝒫}`$ is assured, since it is made of all the fibers of $`\mu `$ at every point of $`𝒫`$. The rest of the proof is like in Proposition 7. So we have the diagram $$\begin{array}{ccccc}\stackrel{~}{𝒫}& \text{}& & & \pi \\ \text{}& \text{}& \text{}& \text{}& \text{}\\ 𝒫& \text{}& 𝒫& \text{}& J^1\pi ^{}\end{array}$$ Bearing in mind Remark 16, we can also state: ###### Corollary 2 Let ($`\stackrel{~}{𝒫},\alpha _{\stackrel{~}{𝒫}})`$ be the couple associated with a given restricted almost-regular Hamiltonian system $`(J^1\pi ^{},𝒫,h_𝒫)`$ by the Proposition 14. Consider the characteristic distribution $`𝒟_{\alpha _{\stackrel{~}{𝒫}}}`$, and let $`\{h\}_{\alpha _{\stackrel{~}{𝒫}}}`$ be the family of local sections of $`\stackrel{~}{\mu }_{\stackrel{~}{𝒫}}`$ such that $`\mathrm{Im}hS`$ are local integral submanifolds of $`𝒟_{\alpha _{\stackrel{~}{𝒫}}}`$. Then, for every $`\stackrel{~}{h}_𝒫^{}\{h\}_{\alpha _{\stackrel{~}{𝒫}}}`$, we have that $`\mathrm{Im}\stackrel{~}{h}_𝒫^{}`$ is locally a level set of a function $`\mathrm{H}_{\stackrel{~}{𝒫}}`$ such that $`\alpha _{\stackrel{~}{𝒫}}|_S=(\mathrm{dH}_{\stackrel{~}{𝒫}})|_S`$, locally. ###### Definition 12 Given a restricted almost-regular Hamiltonian system $`(J^1\pi ^{},𝒫,h_𝒫)`$, let ($`\stackrel{~}{𝒫},\alpha _{\stackrel{~}{𝒫}})`$ be the couple associated with $`(J^1\pi ^{},𝒫,h_𝒫)`$ by the Proposition 14. The triple $`(\pi ,\stackrel{~}{𝒫},\alpha _{\stackrel{~}{𝒫}})`$ will be called the (local) extended almost-regular Hamiltonian system associated with $`(J^1\pi ^{},h)`$. ###### Proposition 15 Let $`\{(J^1\pi ^{},𝒫,h_𝒫)\}`$ be the class of restricted almost-regular Hamiltonian systems associated with an extended almost-regular Hamiltonian system $`(\pi ,\stackrel{~}{𝒫},\alpha _{\stackrel{~}{𝒫}})`$. Consider the submanifolds $`\{S_{h_𝒫}=\mathrm{Im}h_𝒫\}`$, for every Hamiltonian section $`h`$ in this class, and let $`ȷ_{S_{h_𝒫}}:S_{h_𝒫}\stackrel{~}{𝒫}`$ be the natural embeddings. Then the submanifolds $`\{S=\mathrm{Im}h_𝒫\}`$, for every Hamiltonian section $`h_𝒫`$ in this class are premultisymplectomorphic. As a consequence of this, if $`𝒳_{\alpha _{\stackrel{~}{𝒫}}}\text{X}^m(\stackrel{~}{𝒫})`$ is a solution to the equations (20), the multivector fields $`𝒳_{S_{h_𝒫}}\text{X}^m(S_{h_𝒫})`$ such that $`\mathrm{\Lambda }^m(\stackrel{~}{ȷ}_{S_{h_𝒫}})_{}𝒳_{S_{h_𝒫}}=𝒳_{\alpha _{\stackrel{~}{𝒫}}}|_{S_{h_𝒫}}`$, for every submanifold $`S_{h_𝒫}`$ of this family, are related by these presymplectomorphisms. ## 6 Examples ### 6.1 Restricted Hamiltonian system associated with a Lagrangian system A particular but relevant case concerns (first-order) Lagrangian field theories and their Hamiltonian counterparts. In field theory, a Lagrangian system is a couple $`(J^1\pi ,\mathrm{\Omega }_{})`$, where $`J^1\pi `$ is the first-order jet bundle of $`\pi :EM`$, and $`\mathrm{\Omega }_{}\mathrm{\Omega }^{m+1}(J^1\pi )`$ is the Poincaré-Cartan $`(m+1)`$-form associated with the Lagrangian density $``$ describing the system ($``$ is a $`\overline{\pi }^1`$-semibasic $`m`$-form on $`J^1\pi `$, which is written as $`=\mathrm{\pounds }\overline{\pi }^1\eta \mathrm{\pounds }\omega `$, where $`\mathrm{\pounds }\mathrm{C}^{\mathrm{}}(J^1\pi )`$ is the Lagrangian function associated with $``$ and $`\omega `$). The Lagrangian system is regular if $`\mathrm{\Omega }_{}`$ is $`1`$-nondegenerate; elsewhere it is singular. The extended Legendre map associated with $``$, $`\stackrel{~}{}:J^1\pi \pi `$, is defined by $$(\stackrel{~}{}(\overline{y}))(Z_1,\mathrm{},Z_m):=(\mathrm{\Theta }_{})_{\overline{y}}(\overline{Z}_1,\mathrm{},\overline{Z}_m)$$ where $`Z_1,\mathrm{},Z_m\mathrm{T}_{\pi ^1(\overline{y})}E`$, and $`\overline{Z}_1,\mathrm{},\overline{Z}_m\mathrm{T}_{\overline{y}}J^1\pi `$ are such that $`\mathrm{T}_{\overline{y}}\pi ^1\overline{Z}_{\alpha _{\stackrel{~}{𝒫}}}=Z_{\alpha _{\stackrel{~}{𝒫}}}`$. ($`\stackrel{~}{}`$ can also be defined as the “first order vertical Taylor approximation to $`\mathrm{\pounds }`$). We have that $`\stackrel{~}{}^{}\mathrm{\Omega }=\mathrm{\Omega }_{}`$. If $`(x^\alpha ,y^A,v_\alpha ^A)`$ is a natural chart of coordinates in $`J^1\pi `$ (adapted to the bundle structure, and such that $`\omega =\mathrm{d}x^1\mathrm{}\mathrm{d}x^m\mathrm{d}x^m`$) the local expressions of these maps are $$\begin{array}{ccccccc}\stackrel{~}{}^{}x^\nu =x^\nu & ,& \stackrel{~}{}^{}y^A=y^A& ,& \stackrel{~}{}^{}p_A^\nu =\frac{\mathrm{\pounds }}{v_\nu ^A}& ,& \stackrel{~}{}^{}p=\mathrm{\pounds }v_\nu ^A\frac{\mathrm{\pounds }}{v_\nu ^A}\\ ^{}x^\nu =x^\nu & ,& ^{}y^A=y^A& ,& ^{}p_A^\nu =\frac{\mathrm{\pounds }}{v_\nu ^A}& & \end{array}$$ Using the natural projection $`\mu :\pi J^1\pi ^{}`$, we define the restricted Legendre map associated with $``$ as $`:=\mu \stackrel{~}{}`$. Then, $`(J^1\pi ,\mathrm{\Omega }_{})`$ is a regular Lagrangian system if $``$ is a local diffeomorphism (this definition is equivalent to that given above). Elsewhere $`(J^1\pi ,\mathrm{\Omega }_{})`$ is a singular Lagrangian system. As a particular case, $`(J^1\pi ,\mathrm{\Omega }_{})`$ is a hyper-regular Lagrangian system if $``$ is a global diffeomorphism. Finally, a singular Lagrangian system $`(J^1\pi ,\mathrm{\Omega }_{})`$ is almost-regular if: $`𝒫:=(J^1\pi )`$ is a closed submanifold of $`J^1\pi ^{}`$, $``$ is a submersion onto its image, and for every $`\overline{y}J^1\pi `$, the fibers $`^1((\overline{y}))`$ are connected submanifolds of $`J^1E`$. If $`(J^1\pi ,\mathrm{\Omega }_{})`$ is a hyper-regular Lagrangian system, then $`\stackrel{~}{}(J^1\pi )`$ is a 1-codimensional imbedded submanifold of $`\pi `$, which is transverse to the projection $`\mu `$, and is diffeomorphic to $`J^1\pi ^{}`$. This diffeomorphism is $`\mu ^1`$, when $`\mu `$ is restricted to $`\stackrel{~}{}(J^1\pi )`$, and also coincides with the map $`h:=\stackrel{~}{}^1`$, when it is restricted onto its image (which is just $`\stackrel{~}{}(J^1\pi )`$). This map $`h`$ is the Hamiltonian section needed to construct the restricted Hamiltonian system associated with $`(J^1\pi ,\mathrm{\Omega }_{})`$. In other words, the Hamiltonian section $`h`$ is given by the image of the extended Legendre map. Using charts of natural coordinates in $`J^1\pi ^{}`$ and $`\pi `$, we obtain that the local Hamiltonian function $`\mathrm{h}`$ representing this Hamiltonian section is $$\mathrm{h}(x^\nu ,y^A,p_A^\nu )=(^1)^{}\left(v_\nu ^A\frac{\mathrm{\pounds }}{v_\nu ^A}\mathrm{\pounds }\right)=p_A^\nu (^1)^{}v_\nu ^A^1^{}\mathrm{\pounds }$$ Of course, if $`(\pi ,\mathrm{\Omega },\alpha )`$ is any extended Hamiltonian system associated with $`(J^1\pi ^{},h)`$, then $`\stackrel{~}{}(J^1\pi )`$ is an integral submanifold of the characteristic distribution of $`\alpha `$. In an analogous way, if $`(J^1\pi ,\mathrm{\Omega }_{})`$ is an almost-regular Lagrangian system, and the submanifold $`ȷ_{\stackrel{~}{𝒫}}:𝒫J^1\pi ^{}`$ is a fiber bundle over $`E`$ and $`M`$, the $`\mu `$-transverse submanifold $`ȷ:\stackrel{~}{}(J^1\pi )\pi `$ is diffeomorphic to $`𝒫`$. This diffeomorphism $`\stackrel{~}{\mu }:\stackrel{~}{}(J^1\pi )𝒫`$ is just the restriction of the projection $`\mu `$ to $`\stackrel{~}{}(J^1\pi )`$. Then, taking the Hamiltonian section $`h_𝒫:=ȷ\stackrel{~}{\mu }^1`$, we define the Hamilton-Cartan forms $$\mathrm{\Theta }_{h_𝒫}=h_𝒫^{}\mathrm{\Theta }\mathrm{\Omega }_{h_𝒫}=h_𝒫^{}\mathrm{\Omega }$$ which verify that $`_𝒫^{}\mathrm{\Theta }_h^0=\mathrm{\Theta }_{}`$ and $`_𝒫^{}\mathrm{\Omega }_h^0=\mathrm{\Omega }_{}`$ (where $`_𝒫`$ is the restriction map of $``$ onto $`𝒫`$). Once again, this Hamiltonian section $`h_𝒫`$ is given by the image of the extended Legendre map. Then $`(J^1\pi ^{},𝒫,h_𝒫)`$ is the Hamiltonian system associated with the almost-regular Lagrangian system $`(J^1\pi ,\mathrm{\Omega }_{})`$, and we have the following diagram $$\begin{array}{cccc}\text{}& \text{}& \text{}& \text{}\\ & & \text{}& \\ & & M& \end{array}$$ The construction of the (local) extended almost-regular Hamiltonian system associated with $`(J^1\pi ^{},𝒫,h_𝒫)`$ can be made by following the procedure described in section 5.4. Of course, if $`(\pi ,\stackrel{~}{𝒫},\alpha _{\stackrel{~}{𝒫}})`$ is the extended Hamiltonian system associated with $`(J^1\pi ^{},𝒫,h_𝒫)`$, then $`\stackrel{~}{}(J^1\pi )`$ is an integral submanifold of the characteristic distribution of $`\alpha _{\stackrel{~}{𝒫}}`$. ### 6.2 Non-autonomous dynamical systems Another example consists in showing how the so-called extended formalism of time-dependent mechanics (see , , , , ) can be recovered from this more general framework. The starting point consists in giving the configuration bundle, which for a large class of non-autonomous dynamical systems can be taken to be $`\pi :EQ\times `$, where $`Q`$ is a $`n`$-dimensional differentiable manifold endowed with local coordinates $`(q^i)`$, and $``$ has as a global coordinate $`t`$. The extended and restricted momentum phase spaces are $$\pi \mathrm{T}^{}E\mathrm{T}^{}(Q\times )\mathrm{T}^{}Q\times \times ^{},J^1\pi ^{}\mathrm{T}^{}Q\times $$ Then, the following projections can be defined $`pr_1:\pi \mathrm{T}^{}Q`$ , $`\mu :\pi \mathrm{T}^{}Q\times `$ $`pr_2:\pi \times ^{}`$ , $`p:\pi ^{}`$ If $`\mathrm{\Omega }_QZ^2(\mathrm{T}^{}Q)`$ and $`\mathrm{\Omega }_{}Z^2(\times ^{})`$ denote the natural symplectic forms of $`\mathrm{T}^{}Q`$ and $`\times ^{}`$, then the natural symplectic structure of $`\pi `$ is just $$\mathrm{\Omega }=pr_1^{}\mathrm{\Omega }_Q+pr_2^{}\mathrm{\Omega }_{}.$$ Then, we define the so-called extended time-dependent Hamiltonian function $$\mathrm{H}:=\mu ^{}\mathrm{h}+p\mathrm{C}^{\mathrm{}}(\mathrm{T}^{}(Q\times ))$$ where the dynamical information is given by the “time-dependent Hamiltonian function” $`\mathrm{h}\mathrm{C}^{\mathrm{}}(\mathrm{T}^{}Q\times )`$. Now we have that $`(\mathrm{T}^{}(Q\times ),\mathrm{\Omega },\alpha )`$, with $`\alpha =\mathrm{dH}`$, is an extended global Hamiltonian system, and then the equations of motion are $$𝑖(X_\mathrm{H})\mathrm{\Omega }=\mathrm{dH},𝑖(X_\mathrm{H})\mathrm{d}t=1\text{with }X_\mathrm{H}\text{X}(\mathrm{T}^{}(Q\times ))$$ (21) In order to analyze the information given by this equation, we take a local chart of coordinates $`(q^i,p_i,t,p)`$ in $`\mathrm{T}^{}(Q\times )`$, and one can check that the unique solution to these equations is $`X_\mathrm{H}`$ $`=`$ $`{\displaystyle \frac{\mathrm{H}}{p_i}}{\displaystyle \frac{}{q^i}}{\displaystyle \frac{\mathrm{H}}{q^i}}{\displaystyle \frac{}{p_i}}+{\displaystyle \frac{}{t}}{\displaystyle \frac{\mathrm{H}}{t}}{\displaystyle \frac{}{p}}`$ (22) $`=`$ $`{\displaystyle \frac{(\mu ^{}\mathrm{h})}{p_i}}{\displaystyle \frac{}{q^i}}{\displaystyle \frac{(\mu ^{}\mathrm{h})}{q^i}}{\displaystyle \frac{}{p_i}}+{\displaystyle \frac{}{t}}{\displaystyle \frac{(\mu ^{}\mathrm{h})}{t}}{\displaystyle \frac{}{p}}`$ If $`\stackrel{~}{\psi }(t)=(q^i(t),p_i(t),t,p(t))`$ denote the integral curves of this vector field, the last expression leads to the following system of extended Hamiltonian equations $$\frac{d(q^i\stackrel{~}{\psi })}{dt}=\frac{(\mu ^{}\mathrm{h})}{p_i}\stackrel{~}{\psi },\frac{d(p_i\stackrel{~}{\psi })}{dt}=\frac{(\mu ^{}\mathrm{h})}{q^i}\stackrel{~}{\psi },\frac{d(p\stackrel{~}{\psi })}{dt}=\frac{(\mu ^{}\mathrm{h})}{t}\stackrel{~}{\psi }$$ (23) Observe that the last equation corresponds to the last group of equations (17) in the general case of field theories. In fact, using the other Hamilton equations we get $`{\displaystyle \frac{d(p\stackrel{~}{\psi })}{dt}}`$ $`=`$ $`{\displaystyle \frac{(\mu ^{}\mathrm{h}\psi )}{dt}}={\displaystyle \frac{(\mu ^{}\mathrm{h})}{t}}|_{\stackrel{~}{\psi }}{\displaystyle \frac{(\mu ^{}\mathrm{h})}{q^i}}|_{\stackrel{~}{\psi }}{\displaystyle \frac{(q^i\stackrel{~}{\psi })}{t}}{\displaystyle \frac{(\mu ^{}\mathrm{h})}{p_i}}|_{\stackrel{~}{\psi }}{\displaystyle \frac{(p_i\stackrel{~}{\psi })}{t}}`$ $`=`$ $`{\displaystyle \frac{(\mu ^{}\mathrm{h})}{t}}|_{\stackrel{~}{\psi }}{\displaystyle \frac{(\mu ^{}\mathrm{h})}{q^i}}|_{\stackrel{~}{\psi }}{\displaystyle \frac{(\mu ^{}\mathrm{h})}{p_i}}|_{\stackrel{~}{\psi }}+{\displaystyle \frac{(\mu ^{}\mathrm{h})}{p_i}}|_{\stackrel{~}{\psi }}{\displaystyle \frac{(\mu ^{}\mathrm{h})}{q^i}}|_{\stackrel{~}{\psi }}={\displaystyle \frac{(\mu ^{}\mathrm{h})}{q^i}}|_{\stackrel{~}{\psi }}`$ However, as the physical states are the points of $`\mathrm{T}^{}Q\times `$ and not those of $`\mathrm{T}^{}(Q\times )`$, the vector field which gives the real dynamical evolution is not $`X_\mathrm{H}`$, but another one in $`\mathrm{T}^{}Q\times `$ which, as $`X_\mathrm{H}`$ is $`\mu `$-projectable, is just $`\mu _{}X_\mathrm{H}=X_h\text{X}(\mathrm{T}^{}Q\times )`$, that is, in local coordinates $`(q^i,p_i,t)`$ of $`\mathrm{T}^{}Q\times `$, $$X_h=\frac{\mathrm{h}}{p_i}\frac{}{q^i}\frac{\mathrm{h}}{q^i}\frac{}{p_i}+\frac{}{t}$$ (24) Thus, the integral curves $`\psi (t)=(q^i(t),p_i(t),t)`$ of $`X_h`$ are the $`\mu `$-projection of those of $`X_H`$, and they are solutions to the system of Hamilton equations $$\frac{d(q^i\psi )}{dt}=\frac{h}{p_i}\psi ,\frac{d(p_i\psi )}{dt}=\frac{h}{q^i}\psi $$ This result can also be obtained by considering the class of restricted Hamiltonian systems associated with $`(\mathrm{T}^{}(Q\times ),\mathrm{\Omega },\mathrm{dH})`$. In fact, $`\mathrm{T}^{}(Q\times )`$ is foliated by the family of hypersurfaces of $`\mathrm{T}^{}(Q\times )`$ where the extended Hamiltonian function is constant; that is, $$S:=\{𝐩\mathrm{T}^{}(Q\times )\mathrm{H}(𝐩)=r(ctn.)\}$$ which are the integral submanifolds of the characteristic distribution of $`\alpha =\mathrm{dH}`$. Thus, every $`S`$ is defined in $`\mathrm{T}^{}(Q\times )`$ by the constraint $`\zeta :=\mathrm{H}r`$, and the vector field given in (22), which is the solution to (21), is tangent to all of these submanifolds. Then, taking the global Hamiltonian sections $$h:(q^i,p_i,t)(q^i,p_i,t,p=r\mu ^{}\mathrm{h})$$ we can construct the restricted Hamiltonian systems $`(\mathrm{T}^{}Q\times ,h)`$ associated with $`(\mathrm{T}^{}(Q\times ),\mathrm{\Omega },\mathrm{dH})`$. Therefore (24) is the solution to the equations $$𝑖(X_h)\mathrm{\Omega }_h=0,𝑖(X_h)\mathrm{d}t=1\text{with }X_\mathrm{H}\text{X}(\mathrm{T}^{}(Q\times ))$$ where $$\mathrm{\Omega }_h=h^{}\mathrm{\Omega }=\omega _Q\mathrm{dh}\mathrm{d}t\mathrm{\Omega }^2(\mathrm{T}^{}Q\times )$$ The dynamics on each one of these restricted Hamiltonian systems is associated to a given constant value of the extended Hamiltonian. Observe also that, on every submanifol $`S`$, the global coordinate $`p`$ is identified with the physical energy by means of the time-dependent Hamiltonian function $`\mu ^{}\mathrm{h}`$, and hence the last equation (23) shows the known fact that the energy is not conserved on the dynamical trajectories of time-dependent systems. In this way, we have also recovered one of the standard Hamiltonian formalisms for time-dependent systems (see ). ## 7 Conclusions and outlook The usual way of defining Hamiltonian systems in first-order field theory consists in working in the restricted multimomentum bundle $`J^1\pi ^{}`$, which is the natural multimomentum phase space for field theories, but $`J^1\pi ^{}`$ has no a natural multisymplectic structure. Thus, in order to define restricted Hamiltonian systems we use Hamiltonian sections $`h:J^1\pi ^{}\pi `$, which carry the ‘physical information’ and allow us to pull-back the natural multisymplectic structure of $`\pi `$ to $`J^1\pi ^{}`$. In this way we obtain the Hamilton-Cartan form $`\mathrm{\Omega }_h\mathrm{\Omega }^{m+1}(J^1\pi ^{})`$, and then the Hamiltonian field equations can be derived from the Hamilton-Jacobi variational principle. As a consequence, both the geometry and the ‘physical information’ are coupled in the non-canonical multisymplectic form $`\mathrm{\Omega }_h`$. The alternative way that we have introduced consists in working directly in the extended multimomentum bundle $`\pi `$, which is endowed with a canonical multisymplectic structure $`\mathrm{\Omega }\mathrm{\Omega }^{m+1}(\pi )`$. Then we define extended Hamiltonian systems as a triple $`(\pi ,\mathrm{\Omega },\alpha )`$, where $`\alpha Z^1(\pi )`$ is a $`\mu `$-transverse closed form, and the Hamiltonian equation is $`𝑖(𝒳)\mathrm{\Omega }=(1)^{m+1}\alpha `$, with $`𝒳\text{X}^m(\pi )`$. Thus, in these models, the geometry $`\mathrm{\Omega }`$ and the ‘physical information’ $`\alpha `$ are not coupled, and geometric field equations can be expressed in an analogous way to those of mechanical autonomous Hamiltonian systems. The characteristic distribution $`𝒟_\alpha `$ associated with $`\alpha `$, being involutive, has 1-codimensional and $`\mu `$-transverse integrable submanifolds of $`\pi `$, where the sections solution to the field equations are contained. These integrable submanifolds can be locally identified with local sections of the affine bundle $`\mu :\pi J^1\pi ^{}`$. Each one of them allows us to define locally a restricted Hamiltonian system, although all those associated with the same form $`\alpha `$ are, in fact, multisymplectomorphic. The conditions for the existence of global Hamiltonian sections have been also analyzed. Conversely, every restricted Hamiltonian system is associated with an extended Hamiltonian system (at least locally). In addition, the extended Hamiltonian field equations can be obtained from an extended Hamilton-Jacobi variational principle, stated on the set of sections of the bundle $`\overline{\kappa }:\pi M`$, which are integral sections of the characteristic distribution of $`\alpha `$, taking the variations given by the set of the $`\overline{\kappa }`$-vertical vector fields incident to $`\alpha `$. In fact, a part of the local system of differential equations for the critical sections of an extended Hamiltonian system is the same as for the associated restricted Hamiltonian system. Nevertheless, there is another part of the whole system of differential equations which leads to the condition that the critical sections must also be integral submanifolds of the characteristic distribution $`𝒟_\alpha `$. Restricted and extended Hamiltonian systems for submanifolds of $`J^1\pi ^{}`$ and $`\pi `$ (satisfying suitable conditions) have been defined in order to include the almost-regular field theories in this picture. Their properties are analogous to the former case. The extended Hamiltonian formalism has already been used for defining Poisson brackets in field theories . It could provide new insights into some classical problems, such as: reduction of multisymplectic Hamiltonian systems with symmetry, integrability, and quantization of multisymplectic Hamiltonian field theories. ### Acknowledgments We acknowledge the financial support of Ministerio de Educación y Ciencia, projects BFM2002-03493, MTM2004-7832 and MTM2005-04947. We wish to thank Mr. Jeff Palmer for his assistance in preparing the English version of the manuscript.
warning/0506/hep-ph0506325.html
ar5iv
text
# 1. Introduction ## 1. Introduction The wealth of recent data from BES and CLEO has led to a welcome revival of interest in charmonium physics. The data now provide an ideal testing ground for theoretical expectations on the decay mechanisms at work in the $`c\overline{c}`$ system. The conventional picture of a strong three-gluon annihilation of the $`\mathrm{\Psi }^{}`$ runs into more and more difficulties. The so-called $`\rho \pi `$ puzzle and hadronic excess in $`\mathrm{\Psi }^{}`$ decays pose indeed challenging problems . A few years ago, two of us proposed a simple scheme for the decays of the $`J/\mathrm{\Psi }`$ and the $`\mathrm{\Psi }^{}`$. In particular, it was suggested that the $`\mathrm{\Psi }^{}`$ does not significantly annihilate into three gluons. In this note, we update and sharpen the arguments which led to this somewhat unconventional point of view. In our scenario, all non-electromagnetic hadronic decays of the $`\mathrm{\Psi }^{}`$ have a simple and general explanation: survival amplitudes. By this, we mean transition amplitudes from $`\mathrm{\Psi }^{}`$ to lower-lying states which still contain a $`c\overline{c}`$ pair. One original point of this note is the proposal that the exclusive channel $`\mathrm{\Psi }^{}\eta _c+(3\pi )`$ could account for a significant fraction (possibly more than $`1\%`$!) of all $`\mathrm{\Psi }^{}`$ decays. We do urge our experimental colleagues to actively search for this forgotten decay mode. In Section 2 we briefly review and motivate our point of view on $`c\overline{c}`$ annihilation into gluons. For the $`\mathrm{\Psi }^{}`$, survival precedes annihilation! More precisely, the $`c\overline{c}`$ pair survives by spitting out two or three non-perturbative (i.e. with energy much less than 1 Gev) gluons and the lower lying pair then annihilates into three or two perturbative (i.e. with energy $``$ 1 Gev) gluons, depending on the quantum numbers. These 2+3 or 3+2 annihilation scenarii get rid of the so-called hadronic excess problem in $`\mathrm{\Psi }^{}`$ decays. Many experimental tests of these ideas are possible. To lowest order, survival amplitudes are not expected to hadronize in two-body channels. It follows that these decays of the $`\mathrm{\Psi }^{}`$ should result from a direct electromagnetic annihilation of the $`c\overline{c}`$ pair. Tests and predictions of this assertion will be discussed in Section 3. In particular, as already emphasized in our earlier paper, the $`\rho \pi `$ puzzle is then simply solved. To conclude this note we comment very briefly on some other issues in charmonium physics. ## 2. Strong $`c\overline{c}`$ annihilation It is well known that in the $`J/\mathrm{\Psi }`$ $`(1^{})`$ decays the $`c\overline{c}`$ pair mainly annihilates into three perturbative gluons $`(3g)`$ or into a photon. The least massive channel into which the $`3g`$ can materialize is $`\rho \pi `$ which is indeed the strongest observed hadronic two-body decay of the $`J/\mathrm{\Psi }`$. However, there is also a significant $`c\overline{c}`$ survival amplitude namely $`J/\mathrm{\Psi }\eta _c\gamma `$. Despite the cost of emitting a photon, this decay has the same branching ratio as the $`\rho \pi `$ channel. For the $`\mathrm{\Psi }^{}`$ $`(1^{})`$, the survival radiative decays $`\mathrm{\Psi }^{}\gamma +\chi _c`$ or $`\eta _c`$ $`(0^{++}\mathrm{,1}^{++}\mathrm{,2}^{++}`$ or $`0^+)`$ are quite important. Similarly, the dominant strong decay channels have the structure $$\mathrm{\Psi }^{}(2NPg)+(3g).$$ (1) The physical picture is as simple as can be: the excited $`c\overline{c}`$ pair in the $`\mathrm{\Psi }^{}`$ does not annihilate directly but rather in a two-step process. By spitting out two non-perturbative gluons $`(2NPg)`$, it first survives in a lower $`c\overline{c}`$ configuration ($`1^{}`$ or $`1^+`$) which then eventually annihilates into $`3g`$. The decays $$\mathrm{\Psi }^{}(2\pi )J/\mathrm{\Psi }$$ (2a) $$\mathrm{\Psi }^{}\eta J/\mathrm{\Psi }$$ (2b) clearly follow the pattern of Eq. (1). Particularly important from our point of view is the recently observed survival decay $$\mathrm{\Psi }^{}\pi ^0h_c.$$ (3) It is also of the type Eq. (1) where the $`(2NPg)`$, the $`\eta _0`$ in this case, mixes with the $`\pi ^0`$. The observed rate implies that the effective $`\mathrm{\Psi }^{}h_c\eta _0`$ coupling is of the same order as the coupling $`\mathrm{\Psi }^{}J/\mathrm{\Psi }\eta _0`$. At present , the survival radiative decays together with the three on-shell channels (Eqs. (2) and (3)) account for more than 80% of all $`\mathrm{\Psi }^{}`$ decays. The success of the Gell-Mann, Sharp, Wagner off-shell model for the decay $`\omega 3\pi `$, namely $`\omega \pi +\mathrm{`}\mathrm{`}\rho \mathrm{"}`$, has led us to suggest that decay modes, still of the type Eq. (1), $$\begin{array}{cccccccc}& & \mathrm{\Psi }^{}2\pi (0^{++})+\mathrm{`}\mathrm{`}h_c\mathrm{"}(1^+)\hfill & & \text{(4a)}\hfill & & & \\ & & \mathrm{\Psi }^{}\eta (0^+)+\mathrm{`}\mathrm{`}h_c\mathrm{"}(1^+)\hfill & & \text{(4b)}\hfill & & & \end{array}$$ might also be the source of sizeable light hadron decay modes of the $`\mathrm{\Psi }^{}`$. The observed and large $`5\pi `$ hadronic decay of the $`\mathrm{\Psi }^{}`$ could already correspond to the pattern of Eq. (4a) where the $`\mathrm{`}\mathrm{`}h_c\mathrm{"}`$ is only slightly off-shell. It would be nice if, for this decay, experimentalists could identify a two-pion invariant mass with the quantum numbers $`0^{++}`$. There is of course another possibility for a two-step decay pattern $$\mathrm{\Psi }^{}(3NPg)+(2g)$$ (5) where the lower $`c\overline{c}`$ configuration ($`0^+`$ or $`0^{++}`$) annihilates into $`2g`$. The only on-shell channel for this type of decays is $$\mathrm{\Psi }^{}(3\pi )\eta _c$$ (6a) to which one may again add the least off-shell amplitude $$\mathrm{\Psi }^{}3\pi (1^{})+\mathrm{`}\mathrm{`}\chi _{c_0}\mathrm{"}(0^{++}).$$ (6b) Eq. (6a) is an original ingredient of this note. It is a genuine survival amplitude corresponding to the process where the $`c\overline{c}`$ pair in the $`\mathrm{\Psi }^{}`$ falls to a lower configuration ($`\eta _c`$) by radiating three non-perturbative gluons which hadronize in $`3\pi `$. It could easily correspond to $`1\%`$ or more of all $`\mathrm{\Psi }^{}`$ decays. An effective calculation (i.e. at the hadronic level) of this decay amplitude requires some guesswork about couplings which leaves room for considerable uncertainty. Details of these calculations will be presented elsewhere. It is however quite interesting to point out that the dominant hadronic decay mode $`\mathrm{\Psi }^{}7\pi `$ naturally follows from Eq. (6a). We beg experimentalists to search for a $`\eta _c`$ peak in this multi-pion final state. In summary, Eqs. (1) and (5) are our explanation of the so-called hadronic excess in $`\mathrm{\Psi }^{}`$ decays. If true, there appears to be no need whatsoever for an important contribution of direct $`\mathrm{\Psi }^{}`$ annihilation into three gluons. Furthermore, the substitution of one photon for one gluon in Eqs. (1) and (5) allows $$\mathrm{\Psi }^{}(2NPg)+2g+\gamma .$$ (7) This 2+2+1 pattern corresponds to on-shell radiative decays such as $$\mathrm{\Psi }^{}(\pi ^+\pi ^{})\eta _c\gamma $$ (8a) $$\mathrm{\Psi }^{}\eta \eta _c\gamma $$ (8b) which could be larger than the observed $`\mathrm{\Psi }^{}\eta _c\gamma `$ mode. ## 3. Electromagnetic $`c\overline{c}`$ annihilation Whether the $`\mathrm{\Psi }^{}`$ decays following the 2+3 (Eq. (1)) or 3+2 (Eq. (5)) pattern, it seems intuitively difficult to end up with a light hadronic two-body channel. This brings us to the suggestion that these channels are of electromagnetic origin, namely they follow from the direct hadronization of a virtual photon. If such is the case, the $`e^+e^{}\gamma ^{}`$ hadrons continuum should be consistently substracted for all two-body branching ratios. In the $`SU(3)`$ limit for hadrons, a well-known consequence of photon hadronization is that the ratio of branching ratios into neutral and charged strange states is expected to be 4 (for $`d`$-coupling). The recent data on $`\mathrm{\Psi }^{}K^{}\overline{K}`$ agree very well with this expectation. Moreover, a striking difference between $`J/\mathrm{\Psi }`$ and $`\mathrm{\Psi }^{}`$ decay modes is observed in the $`3\pi `$ channel. For the $`J/\mathrm{\Psi }`$, the $`\rho `$(770) almost saturates the two-pion invariant mass, while for the $`\mathrm{\Psi }^{}`$ it is the $`\rho `$(2150) which seems to dominate. Such a strong suppression of the low-lying vector state contributions is not surprising in a high-energy electromagnetic process. These observations considerably strenghten the argument that $`\mathrm{\Psi }^{}VP`$ is dominantly an electromagnetic process: the so-called $`\rho \pi `$ puzzle is solved. Physically, the $`1^+0^+`$ channel $`b_1\pi `$ is even more interesting. For the moment , it is the largest light hadronic two-body decay of the $`\mathrm{\Psi }^{}`$ but still of the same order as $`\mathrm{\Psi }^{}2(\pi ^+\pi ^{})`$ which is obviously of electromagnetic nature. If the $`b_1\pi `$ channel comes from the hadronization of a photon, then both $`\mathrm{\Psi }^{}b_1`$(1235)$`\eta `$ and $`\mathrm{\Psi }^{}h_1`$(1170)$`\pi ^0`$ should have branching ratios of the order of $`10^3`$. These processes are certainly welcome to saturate the theoretically well-known $`\mathrm{\Psi }^{}\gamma ^{}`$ light hadrons branching ratio $`(1.6\%)`$. Moreover, they lead to the surprising prediction that $$\text{Br}(J/\mathrm{\Psi }h_1\pi ^0)\text{Br}(J/\mathrm{\Psi }b_1\eta )1\%$$ (9) which are of the same order as the measured $`J/\mathrm{\Psi }\rho \pi `$ branching ratio. ## 4. Comments and conclusion The main point of this note has been to reemphasize that there is no experimental necessity for a direct strong annihilation of the $`\mathrm{\Psi }^{}`$ into $`3g`$. Why is this annihilation process suppressed? We can only repeat the argument given earlier : the putative (or theoretical) $`c\overline{c}`$ states $`(n^{2S+1}L_J)`$ are one thing, the physical states are quite another! With strong annihilation of the $`1^{}`$ ground state and its first “radial excitation", mixing is expected. The QCD dynamics may be such that the physical states, presumably mixtures of the putative ones, are so built up that one of them strongly annihilates into three perturbative gluons while the other does not. Mixing of the 1 $`{}_{}{}^{3}S_{1}^{}`$ and 2 $`{}_{}{}^{3}S_{1}^{}`$ states via three perturbative gluons has little effect on the charmonium mass spectrum, but may be crucial for the decay pattern. If this explanation is correct, one may wonder about the decay patterns of the $`0^+`$ states below the open charm threshold. For the $`\eta _c^{}`$, we do expect survival to be significantly more important than a direct two-gluon annihilation. Another comment concerns the 12% rule: we do not see any reason for this rule to be valid. Contrary to the electromagnetic annihilation of the $`c\overline{c}`$ into a photon which is a pointlike process, neither the $`J/\mathrm{\Psi }`$ annihilation into $`3g`$ nor the 2+3 or 3+2 patterns for the $`\mathrm{\Psi }^{}`$ are of the same nature except, possibly, in the $`m_c\mathrm{}`$ limit, but then sizeable corrections are to be expected. To conclude let us repeat that the main points of this short note have been: 1. to revive and make more precise a very simple picture of all strong and radiative decay modes of the $`\mathrm{\Psi }^{}`$; 2. to infer that two-body decays of $`\mathrm{\Psi }^{}`$ into light hadrons are of electromagnetic origin. These simultaneously solve the so-called hadronic excess and $`\rho \pi `$ puzzle, respectively. Elegant as this may seem, experimental confirmation is still required. The explicit identification of the decays $`\mathrm{\Psi }^{}\eta _c(3\pi ),\eta _c(2\pi )\gamma `$ and a measurement of the branching ratios for $`J/\mathrm{\Psi }h_1\pi ^0,b_1\eta `$ would be important steps in this direction. ## Acknowledgements This work was supported by the Belgian Federal Office for Scientific, Technical and Cultural Affairs through the Interuniversity Attraction Pole P5/27.
warning/0506/hep-ph0506218.html
ar5iv
text
# Distorting the Hump-backed Plateau of Jets with Dense QCD Matter ## Abstract The hump-backed plateau of the single inclusive distribution of hadrons inside a jet provides a standard test of the interplay between probabilistic parton splitting and quantum coherence in QCD. The medium-induced modification of this QCD radiation physics is expected to give access to the properties of the dense medium produced in relativistic heavy ion collisions. Here, we introduce a formulation of medium-induced parton energy loss, which treats all leading and subleading parton branchings equally, and which – for showering in the vacuum – accounts for the observed distribution of soft jet fragments. We show that the strong suppression of single inclusive hadron spectra measured in Au-Au collisions at the Relativistic Heavy Ion Collider (RHIC) implies a characteristic distortion of the hump-backed plateau; we determine, as a function of jet energy, to what extent the soft jet fragments can be measured above some momentum cut. Our study further indicates that the approximate $`p_T`$-independence of the measured nuclear modification factor does not exclude a significant $`Q^2`$-dependence of parton energy loss. Over the last five years, experiments at RHIC have established a phenomenon of strong high-$`p_T`$ hadron suppression Adcox:2004mh ; Adams:2005dq , which supports the picture that high-$`p_T`$ partons produced in the dense matter of a nuclear collision suffer a significant energy degradation prior to hadronization in the vacuum Jacobs:2004qv . The microscopic dynamics conjectured to underly high-$`p_T`$ hadron suppression is medium-induced gluon radiation Baier:2000mf ; Kovner:2003zj ; Gyulassy:2003mc , i.e., a characteristic medium-induced distortion of the standard QCD radiation pattern tested extensively by jet measurements in high energy $`e^+e^{}`$ and $`pp(p\overline{p})`$ collisions. Modeling this effect accounts for the measured single-inclusive spectra and leading back-to-back hadron correlations Jacobs:2004qv ; Gyulassy:2003mc ; Dainese:2004te ; Eskola:2004cr . However, existing models indicate only rough qualitative features of subleading jet fragments such as their broadening and softening due to medium-effects. This is so mainly, since medium-induced parton splitting is included for the leading partons only, but not consistently for the subleading splitting processes Baier:2001yt ; Gyulassy:2003mc , and since energy momentum conservation is taken into account globally, but not ensured locally for each parton splitting Gyulassy:2003mc ; Eskola:2004cr . Here, we propose a formulation which overcomes these limitations. Subleading jet fragments are known to provide many fundamental tests of QCD radiation physics. In particular, for soft particle momentum fractions $`x=p/E_{\mathrm{jet}}`$ inside a quark- or gluon- ($`i=q,g`$) initiated jet of energy and virtuality $`QE_{\mathrm{jet}}`$, the single inclusive distribution $`D_i(x,Q^2)`$ is dominated by multiparton destructive interference, and thus tests quantitatively the understanding of QCD coherence Mueller:1982cq ; Dokshitzer:1988bq . Remarkably, to double and single logarithmic accuracy in $`\xi \mathrm{ln}\left[1/x\right]`$ and $`\tau \mathrm{ln}\left[Q/\mathrm{\Lambda }_{\mathrm{eff}}\right]`$, $`\mathrm{\Lambda }_{\mathrm{eff}}=O(\mathrm{\Lambda }_{\mathrm{QCD}})`$, the effects of this destructive quantum interference can be accounted for by an angular ordering prescription of a probabilistic parton cascade with leading order (LO) splitting functions. The so-called modified leading logarithmic approximation (MLLA) resums these effects and accounts for the large $`\sqrt{\alpha _s}`$ next-to-leading corrections of $`D_i(x,Q^2)`$ Bassetto:1984ik ; Mueller:1982cq ; Dokshitzer:1988bq . The MLLA leads to an evolution equation for the $`\nu `$-th Mellin moments $`M_i(\nu ,\tau )=_0^{\mathrm{}}𝑑\xi e^{\nu \xi }xD_i(x,Q^2)`$ Bassetto:1984ik ; Dokshitzer:1988bq ; Fong:1990nt , $`{\displaystyle \frac{}{\tau }}\left(\begin{array}{c}M_q(\nu ,\tau )\\ M_g(\nu ,\tau )\end{array}\right)`$ $`=`$ $`\left[\begin{array}{cc}\mathrm{\Phi }_{qq}\left(\nu +\frac{}{\tau }\right)& \mathrm{\Phi }_{qg}\left(\nu +\frac{}{\tau }\right)\\ \mathrm{\Phi }_{gq}\left(\nu +\frac{}{\tau }\right)& \mathrm{\Phi }_{gg}\left(\nu +\frac{}{\tau }\right)\end{array}\right]`$ (8) $`\times {\displaystyle \frac{\alpha _s(\tau )}{2\pi }}\left(\begin{array}{c}M_q(\nu ,\tau )\\ M_g(\nu ,\tau )\end{array}\right).`$ Here, $`\mathrm{\Phi }_{ij}`$ denote combinations of particular moments of leading order splitting functions, for example $`\mathrm{\Phi }_{qq}\left(\nu \right)`$ $`=`$ $`2{\displaystyle _0^1}𝑑zP_{qq}(z)\left(z^\nu 1\right).`$ (9) The shift $`\left(\nu +\frac{}{\tau }\right)`$ in (8) accounts for angular ordering. For a parton fragmentation which starts at high initial scale $`\tau `$ and ends at some hadronic scale $`\tau _0`$, the solution of (8) has to fulfill the initial conditions $`M(x,\tau _0)=\delta (1x)`$ and $`\frac{}{\tau }M(x,\tau =\tau _0)=0`$, since the parton must not evolve if produced at the hadronic scale. The lowest Mellin moments $`\nu 0`$ determine the main characteristics of $`D_i(x,Q^2)`$. For an approximate solution of (8), one can thus expand the matrix in (8) to next-to-leading order in $`\left(\nu +\frac{}{\tau }\right)`$ and diagonalize it. Its eigenvalue with leading $`1/\left(\nu +\frac{}{\tau }\right)`$-term yields a differential equation of the confluent hypergeometric type Dokshitzer:1988bq . This leads to an analytic expression for $`D(x,Q^2)`$, whose shape does not distinguish between quark and gluon parents, since the multiplicity is dominated in both cases by gluon branching. For the hadronic multiplicity distribution $`dN^h/d\xi `$, one assumes that at the scale $`\tau _0`$, a parton is mapped locally onto a hadron with proportionality factor $`K^hO(1)`$ (”local parton hadron duality”, LPHD) $$\frac{dN^h}{d\xi }=K^hD\left(x,\tau =\mathrm{ln}\left[\frac{Q=E}{\mathrm{\Lambda }_{\mathrm{eff}}}\right]\right).$$ (10) Comparisons of (10) to data have been performed repeatedly Fong:1990nt ; Dokshitzer:1992jv ; Braunschweig:1990yd ; Abbiendi:2002mj ; Acosta:2002gg over a logarithmically wide kinematic regime $`7<E_{\mathrm{jet}}<150`$ GeV in both $`e^+e^{}`$ and $`pp/p\overline{p}`$ collisions. To illustrate the degree of agreement, we reproduce in Fig. 1 two sets of data Braunschweig:1990yd ; Abbiendi:2002mj together with the curves obtained from (10). The parameters $`K^h`$ and $`\mathrm{\Lambda }_{\mathrm{eff}}`$ entering (10) were chosen as in Refs. Braunschweig:1990yd ; Abbiendi:2002mj , $`\mathrm{\Lambda }_{\mathrm{eff}}=254`$ MeV, $`K^h=1.15`$ for $`E_{\mathrm{jet}}=100`$ GeV, $`K^h=1.46`$ for $`E_{\mathrm{jet}}=7`$ GeV. Following Ref. Abbiendi:2002mj , we use $`N_f=3`$. From Fig. 1, we conclude that Eq.(10) accounts reasonably well for the jet multiplicity distribution in the kinematic range accessible in heavy ion collisions at RHIC ($`E_{\mathrm{jet}}10`$ GeV) and at the LHC ($`E_{\mathrm{jet}}100`$ GeV). Corrections not included in (10) are of relative order $`1/\tau `$, which at face value corresponds to a 30% (15%) uncertainty at typical RHIC (LHC) jet energies. Also, the MLLA resums large $`\xi `$, $`\tau \xi `$, but is expected to be less accurate for hard jet fragments, where other improvements are currently sought for Albino:2005gg . Thus, the agreement of (10) to data for the entire $`\xi `$-range is surprisingly good. At least from a pragmatic point of view, (10) can serve as a baseline on top of which one can search for medium effects. The multiplicity distribution $`dN^h/d\xi `$ is dominated by soft gluon bremsstrahlung, $`dI^{\mathrm{vac}}C_R\frac{\alpha _s(k_T^2)}{\pi }\frac{dk_T^2}{k_T^2}\frac{d\omega }{\omega }`$, $`\omega =zE_{\mathrm{jet}}`$, which is described by the singular parts $`\frac{1}{z}`$, $`\frac{1}{(1z)}`$ of the QCD splitting functions entering (9). They determine the leading $`\frac{1}{\nu }`$-terms of the evolution matrix in (8). Remarkably, calculations of the additional medium-induced radiation indicate that $`\omega \frac{dI^{\mathrm{med}}}{d\omega }`$ is $`\frac{1}{\sqrt{\omega }}`$ if the medium is modeled by soft multiple momentum transfers Baier:1996sk ; Salgado:2003gb , and $`\frac{1}{\omega }`$ if the medium is modeled by a single hard momentum transfer Gyulassy:2003mc ; Salgado:2003gb . Thus, parametrically, the additional medium-dependent contributions to the gluon bremsstrahlung are more singular than $`dI^{\mathrm{vac}}`$ for small $`\omega `$ and may thus be expected to dominate the multiplicity distribution (10). However, destructive interference due to finite in-medium path length is known to regulate the soft $`\omega `$-divergence Salgado:2003gb . For the relevant range of soft $`\omega `$, this may be modeled as $`\omega \frac{dI^{\mathrm{med}}}{d\omega }f_{\mathrm{med}}=\mathrm{const}`$. A medium-induced gluon bremsstrahlung spectrum, consistent with this ansatz, was also found in Guo:2000nz . This suggests that medium effects enter (10) by enhancing the singular parts of all LO splitting functions $`P_{gg}`$, $`P_{qg}`$, $`P_{qq}`$ by the same factor $`\left(1+f_{\mathrm{med}}\right)`$, such that for example $$P_{qq}(z)=C_F\left(\frac{2\left(1+f_{\mathrm{med}}\right)}{(1z)_+}(1+z)\right).$$ (11) We do not modify the non-singular subleading terms. On general grounds, one expects that medium-induced rescattering is a nuclear enhanced higher-twist contribution ($`f_{\mathrm{med}}\frac{L}{Q^2}`$Luo:1993ui . This means that it is subleading in an expansion in $`Q^2`$, while being enhanced compared to other higher twist contributions by a factor proportional to the geometrical extension $`L`$ of the target. A $`1/Q^2`$-dependence of $`f_{\mathrm{med}}`$ is also suggested by the following heuristic argument. A hard parton of virtuality $`Q`$ has a lifetime $`1/Q`$ in its own rest frame, and thus a lifetime (in-medium path length) $`t=\frac{1}{Q}\frac{E}{Q}`$ before it branches in the rest frame of the dense matter through which it propagates. Medium effects on a parton in between two branching processes should grow proportional to (some power of) the in-medium path length and thus $`1/Q^2`$ or higher powers thereof. In contrast, jet quenching models Jacobs:2004qv ; Gyulassy:2003mc ; Dainese:2004te ; Eskola:2004cr reproduce inclusive hadron spectra in Au-Au collisions at RHIC by supplementing the standard QCD LO factorized formalism with the probability $`P(\mathrm{\Delta }E)`$ that the produced partons radiate an energy $`\mathrm{\Delta }E`$ due to medium effects prior to hadronization in the vacuum Baier:2001yt $`P(\mathrm{\Delta }E)`$ $`=`$ $`{\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{1}{n!}}\left[{\displaystyle \underset{i=1}{\overset{n}{}}}{\displaystyle 𝑑\omega _i\frac{dI^{\mathrm{med}}(\omega _i)}{d\omega }}\right]\delta \left(\mathrm{\Delta }E{\displaystyle \underset{i=1}{\overset{n}{}}}\omega _i\right)`$ (12) $`\times \mathrm{exp}\left[{\displaystyle _0^{\mathrm{}}}𝑑\omega {\displaystyle \frac{dI^{\mathrm{med}}}{d\omega }}\right].`$ This formula is based on a probabilistic iteration of medium-modified parton splittings, but does not keep track of virtuality or angular ordering. The $`k_T`$-integrated medium-induced contribution $`dI^{\mathrm{med}}`$ is treated on an equal footing with LO vacuum splitting functions. In this sense, the medium-modified fragmentation function $`D_{h/q}^{(\mathrm{med})}(x,Q^2)=_0^1𝑑ϵEP(\mathrm{\Delta }E)\frac{1}{1ϵ}D_{h/q}(\frac{x}{1ϵ},Q^2)`$, $`ϵ=\mathrm{\Delta }E/E`$, entering jet quenching models Jacobs:2004qv ; Gyulassy:2003mc ; Dainese:2004te ; Eskola:2004cr , amounts to a medium-induced $`Q^2`$-independent modification of parton fragmentation. The single inclusive distribution $`D(x,Q^2)`$, supplemented by LPHD, is a fragmentation function. Single inclusive hadron spectra, whose parent partons show a power law spectrum $`1/p_T^{n(p_T)}`$, test $`D(x,Q^2)`$ in the range, in which $`x^{n(p_T)2}D(x,Q^2)`$ has significant support. However, for large $`x`$, the accuracy of the MLLA result for $`D(x,Q^2)`$ becomes questionable Albino:2005gg ; Fong:1990nt . To understand to what extent the MLLA result may still be used from a practical point of view, we have compared it to the KKP parametrization Kniehl:2000fe of fragmentation functions. In the range of $`Q^2`$ and $`x`$ relevant for single inclusive spectra ($`0.4<x<0.9`$), we observe that the KKP and MLLA fragmentation functions both drop by 2 orders of magnitude. They do show somewhat different shapes but – after adjusting the overall normalization – they differ for all $`x`$-values by maximally $`30`$% or significantly less (data not shown). For the nuclear modification factor $`R_{AA}`$, which is the ratio of modified and unmodified single inclusive hadron spectra, and which does not depend on the overall normalization of $`D(x,Q^2)`$, this is a relatively small uncertainty, if one aims at characterizing a factor 5 suppression. We thus conclude that the MLLA fragmentation function obtained from (8) can be used to calculate $`R_{AA}`$. To determine $`R_{AA}`$, we have parametrized the partonic $`p_T`$-spectrum at RHIC energies $`\sqrt{s_{NN}}=200`$ GeV by a power law $`1/p_T^{n(p_T)}`$, $`n(p_T)=7+0.003p_T^2/\mathrm{GeV}^2`$, which accounts for kinematic boundary effects at $`p_TO(\sqrt{s_{NN}})`$. Single inclusive hadron spectra and $`R_{AA}`$ are calculated by convoluting this spectrum with $`D(x,Q^2)`$. Medium effects are included through the factor $`f_{\mathrm{med}}`$ in the singular parts of all LO parton splitting functions, see Eq.(11). As seen in Fig. 2, the choice $`f_{\mathrm{med}}0.6÷0.8`$ reproduces the size of the suppression of $`R_{AA}0.2`$ in central Au-Au collisions at RHIC Adler:2003ii . Jet quenching models based on (12) yield a slightly increasing $`p_T`$-dependence of $`R_{AA}(p_T)`$ for a power law $`n(p_T)=n`$, and a rather flat dependence if trigger bias effects due to the $`p_T`$-dependence of $`n(p_T)`$ are included Dainese:2004te ; Eskola:2004cr . In contrast, the MLLA result for $`R_{AA}(p_T)`$ decreases with increasing $`p_T`$, and this tendency is even more pronounced for a realistic shape of the underlying partonic spectrum, see Fig. 2. The reason is that in the MLLA, parent partons of higher $`p_T`$ have higher initial virtuality $`Qp_T`$, and undergo more medium-induced splittings; this results in a smaller value of $`R_{AA}`$. A proper treatment of nuclear geometry may affect quantitative aspects of Fig. 2, but is unlikely to change this qualitative observation. Hence, the observed flat $`p_T`$-dependence of $`R_{AA}(p_T)`$, one may require (in accordance with the arguments given above) a non-vanishing $`Q^2`$-dependence of $`f_{\mathrm{med}}`$, which would reduce medium-effects on high-$`p_T`$ ($`p_TQ`$) partons. Motivated by this observation, we have attempted to solve Eq.(8) for a non-trivial $`Q^2`$-dependence of the medium-enhancement $`f_{\mathrm{med}}`$. We did not find an analytical solution. However, in the absence of medium effects, the analytical solution (10) is reproduced by Monte Carlo (MC) parton showers based on angular ordering Fong:1990nt . This remains true for non-vanishing $`f_{\mathrm{med}}`$. The present study can serve to check future MC showers implementing (11), and it can be extended in MC studies to include a non-trivial $`Q^2`$-dependence of $`f_{\mathrm{med}}`$. We plan such a MC study, mainly to establish to what extent the approximate $`p_T`$-independence of $`R_{AA}`$ up to $`p_T10`$ GeV allows for a significant $`Q^2`$-dependence of parton energy loss. The question of whether and on what scale these effects are $`1/Q^2`$-suppressed is of obvious importance for heavy ion collisions at the LHC, where medium-modified parton fragmentation can be tested in a logarithmically wide $`Q^2`$-range. What is the distortion of the longitudinal jet multiplicity distribution (10), consistent with the observed factor $`5`$ suppression of $`R_{AA}`$? In contrast to calculations based on (12), the medium-enhanced parton splitting introduced via MLLA conserves energy-momentum exactly at each branching, it treats all secondary branchings of softer gluons equally, and it continues all branchings down to the same hadronic scale. This makes it a qualitatively improved tool for the calculation of longitudinal multiplicity distributions, since it matters obviously for $`dN/d\xi `$ whether one gluon is radiated into the bin $`\xi =\mathrm{ln}\left[E_{\mathrm{jet}}/p_{\mathrm{gluon}}\right]`$, or – after further splitting $`gg(z)g(1z)`$ – two gluons with momentum fractions $`z`$ and $`(1z)`$ into bins $`\xi +\mathrm{ln}\left[1/(1z)\right]`$, $`\xi +\mathrm{ln}\left[1/z\right]`$, respectively. We have calculated $`dN/d\xi `$ for a medium-enhanced parton splitting $`f_{\mathrm{med}}=0.8`$ consistent with $`R_{AA}=0.2`$. Results for jet energies relevant at RHIC and at the LHC are shown in Fig. 1. In general, the multiplicity at large momentum fractions (small $`\xi `$) is reduced and the corresponding energy is redistributed into the soft part of the distribution. The maximum of the multiplicity distribution also shifts to a softer value, but this shift is subleading in $`\sqrt{\alpha _s}`$, $`\xi _{\mathrm{max}}/\tau =\frac{1}{2}+a_{\mathrm{med}}\sqrt{\frac{\alpha _s(\tau )}{32N_c\pi }}`$, where $`a_{\mathrm{med}}=\frac{1}{3}\left(11+12f_{\mathrm{med}}\right)N_c+\frac{2}{3}\frac{N_f}{N_c^2}`$. Many experimental characterizations of the medium-modified internal jet structure in heavy ion collisions at RHIC and at the LHC require a soft momentum cut $`p_T^{\mathrm{cut}}`$ to control effects of the high multiplicity background. Can one observe the increase in soft multiplicity shown in Fig. 1, if such a soft background cut $`p_T^{\mathrm{cut}}`$ is applied? To address this issue, we have calculated from (10) the total hadronic multiplicity $`N^h(p_T>p_T^{\mathrm{cut}})`$ above $`p_T^{\mathrm{cut}}`$. As seen in Fig. 3, the medium-enhanced component of soft multiplicity lies below a critical transverse momentum cut $`p_{T,\mathrm{crit}}^{\mathrm{cut}}`$ which increases significantly with $`E_{\mathrm{jet}}`$. For a typical hard jet at RHIC ($`E_{\mathrm{jet}}=15`$ GeV), the additional soft jet multiplicity lies buried in the soft background, $`p_{T,\mathrm{crit}}^{\mathrm{cut}}1.5`$ GeV. For $`E_{\mathrm{jet}}=100`$ GeV, accessible at the LHC, $`p_{T,\mathrm{crit}}^{\mathrm{cut}}4`$ GeV lies well above a cut which depletes the background multiplicity by a factor 10. For $`E_{\mathrm{jet}}=200`$ GeV, we find $`p_{T,\mathrm{crit}}^{\mathrm{cut}}7`$ GeV. The associated total jet multiplicity $`N^h(p_T>p_{T,\mathrm{crit}}^{\mathrm{cut}})`$ for these jet energies rises with $`E_{\mathrm{jet}}`$ from $`4`$, to $`7`$. Fig. 3 indicates a qualitative advantage in extending jet measurements in an LHC heavy ion run near design luminosity to significantly above $`E_{\mathrm{jet}}100`$ GeV, where a sizeable kinematic range $`2÷3\mathrm{GeV}<p_T<p_{T,\mathrm{crit}}^{\mathrm{cut}}7`$ GeV becomes accessible. This may allow a detailed characterization of the enhanced medium-induced radiation above the soft background. In general, the formulation of parton energy loss within the MLLA formalism allows one to address several fundamental questions, that remain untouched by recent model studies of jet quenching. This concerns in particular the important issue of the $`Q^2`$-dependence of parton energy loss discussed above. Moreover, the use of a probabilistic formulation based on angular ordering can also be viewed as a test of the unproven assumption that the medium-induced destructive interference of multi-parton emission can indeed be accounted for by angular ordering, in close similarity to gluon radiation in the vacuum. We finally note that the formalism introduced here is not limited to a discussion of the hump-backed plateau: e.g. it can be applied to the calculation of two-particle correlations within jets, which have been studied in the absence of medium effects Bassetto:1984ik ; Fong:1990nt . It may also apply to transverse jet broadening Bassetto:1984ik , which for a $`Q^2`$-dependent $`f_{\mathrm{med}}`$ may be significantly reduced since large angle emission would remain essentially unmodified by the medium. We plan to address these open questions in future work. We thank S. Catani, G. Sterman and B. Webber for helpful discussions.
warning/0506/quant-ph0506062.html
ar5iv
text
# Determinism in the one-way model ## I Introduction The recent one-way quantum computing model RB01 ; RB02 ; mqqcs has already drawn considerable attention, because it suggests different physical realizations of quantum computing Nielsen04 ; CMJ04 ; BR04 ; TPKV04 ; TPKV05 ; nature05 ; KPA05 ; BES05 ; CCWD05 . However, whether this fundamentally different model may also suggest new insights in quantum information processing still stands as an open question. Computation in this model, consists of a first phase of preparation and entanglement, followed by 1-qubit measurements and a final round of corrections. Making measurements an integral part of computation will in general induce non-deterministic behaviors. To counter this, both measurements and corrections are allowed to depend on the outcomes of previous measurements. This mechanism of feed-forwarding classical observations is known to be a necessary requirement for the model to be universal DKP04 . Whether and how a given pattern can be controlled so as to obtain a globally deterministic behavior is the question we address in this paper. A variety of methods for constructing measurement patterns have been already proposed mqqcs ; graphstates ; CLN04 that guarantee determinism by construction. We introduce a direct condition on open graph states (graph states with inputs and outputs) which guarantees a strong form of deterministic behavior for a class of one-way measurement patterns defined over them. Remarkably, our condition bears only on the geometric structure of the entangled graph states. This condition singles out a class of patterns with flow, which is stable under sequential and parallel compositions and is large enough to realize all unitary and unitary embedding maps. Patterns with flow have interesting additional properties. First, they are uniformly deterministic, in the sense that no matter what the measurements angles are, the obtained set of corrections, which depends only on the underlying geometry, will make the global behavior deterministic. Second, all computation branches have equal probabilities, which means in particular these probabilities are independent of the inputs, and as a consequence, one can show that all such patterns implement unitary embeddings. Third, a more restricted class of patterns having both flow and reverse flow supports an operation of adjunction, corresponding to time-reversal of unitary operations. This smaller class implements all and only unitary transformations. Moreover, for open graph states with flow, one can derive a direct procedure for realization of unitaries as measurements patterns BDK06 . ## II Measurement Patterns We briefly recall the definition of measurement patterns and various notions of determinism. More detailed introductions can be found in Nielsen05 ; Jozsa05 ; BB06 . In this paper, we will employ an algebraic approach called, the *Measurement Calculus* DKP04 . Computations in a pattern involve a combination of 1-qubit preparations $`N_i^\alpha `$, 2-qubit entanglement operators $`E_{ij}:=Z_{ij}`$ (controlled-$`Z`$), 1-qubit measurements $`M_i^\alpha `$, and 1-qubit Pauli corrections $`X_i`$, $`Z_i`$, where $`i`$, $`j`$ represent the qubits on which each of these operations apply, and $`\alpha `$ is a parameter in $`[0,2\pi )`$. Preparation $`N_i^\alpha `$ prepares qubit $`i`$ in state $`|+_\alpha _i`$, where $`|\pm _\alpha `$ stand for $`\frac{1}{\sqrt{2}}(|0\pm e^{i\alpha }|1)`$. Measurement $`M_i^\alpha `$ is defined by orthogonal projections $`|\pm _\alpha \pm _\alpha |_i`$, applied at qubit $`i`$, with the convention that $`|+_\alpha +_\alpha |_i`$ corresponds to the outcome $`0`$, while $`|_\alpha _\alpha |_i`$ corresponds to $`1`$. Note that we consider here only destructive measurements, i.e. a projection $`|\psi \psi |`$ is always followed by a trace out operator and hence we might write it as $`\psi |`$. Qubits are measured at most once, therefore we may represent unambiguously the outcome of the measurement done at qubit $`j`$ by $`s_j`$. Dependent corrections, used to control non-determinism, will be written $`X_i^{s_j}`$ and $`Z_i^{s_j}`$, with $`X_i^0=Z_i^0=I`$, $`X_i^1=X_i`$, and $`Z_i^1=Z_i`$. A *measurement pattern*, or simply a pattern, is defined by the choice of $`V`$ a finite set of qubits, two possibly overlapping subsets $`I`$ and $`O`$ determining the pattern inputs and outputs, and a finite sequence of commands acting on $`V`$. Such a pattern is said to be *runnable* if it satisfies the following: (R0) no command depends on an outcome not yet measured, (R1) no command acts on a qubit already measured or not yet prepared (except preparation commands), and (R2) a qubit $`i`$ is measured (prepared) if and only if $`i`$ is not an output (input). Write $`_I`$ ($`_O`$) for the Hilbert space spanned by the inputs (outputs). The run of a runnable pattern consists simply in executing each command in sequence. If $`n`$ is the number of measurements (which by (R2) is also the number of non outputs) then the run may follow $`2^n`$ different branches. Each branch is associated with a unique binary string $`𝐬`$ of length $`n`$, representing the classical outcomes of the measurements along that branch, and a unique *branch map* $`A_𝐬`$ representing the linear transformation from $`_I`$ to $`_O`$ along that branch. Branch maps decompose as $`A_𝐬=C_𝐬\mathrm{\Pi }_𝐬U`$, where $`C_𝐬`$ is a unitary map over $`_O`$ collecting all corrections on outputs, $`\mathrm{\Pi }_𝐬`$ is a projection from $`_V`$ to $`_O`$ representing the particular measurements performed along the branch, and $`U`$ is a unitary embedding from $`_I`$ to $`_V`$ collecting the branch preparations, and entanglements. Therefore $$\begin{array}{ccccccccccccccc}_𝐬A_𝐬^{}A_𝐬=_𝐬U^{}\mathrm{\Pi }_𝐬U=I\hfill & & & & & & & & & & & & & & \end{array}$$ and $`T(\rho ):=_𝐬A_𝐬\rho A_𝐬^{}`$ is a trace-preserving completely-positive map (cptp-map), explicitly given as a Kraus decomposition. One says that the pattern *realizes* $`T`$. A pattern is said to be *deterministic* if it realizes a cptp-map that sends pure states to pure states. This is equivalent to saying that branch maps are proportional, that is to say, for all $`q_I`$ and all $`𝐬_1`$, $`𝐬_2_2^n`$, $`A_{𝐬_1}(q)`$ and $`A_{𝐬_2}(q)`$ differ only up to a scalar. A pattern is said to be *strongly deterministic* when branch maps are equal, i.e., for all $`𝐬_1`$, $`𝐬_2_2^n`$, $`A_{𝐬_1}=A_{𝐬_2}`$. A pattern is said to be *uniformly deterministic* if it is deterministic for all values of its measurement angles. ###### Lemma 1 Strongly deterministic patterns realize unitary embedding maps. Proof. If a pattern is strongly deterministic and realizes the map $`T`$ then $$\begin{array}{ccccccccccccccc}T(\rho )=A\rho A^{}\hfill & & & & & & & & & & & & & & \end{array}$$ with $`A:=2^{n/2}A_𝐬`$, and $`A`$ must be a unitary embedding, because $`_𝐬A_𝐬^{}A_𝐬=A^{}A=I`$. In such cases, one says that the pattern realizes the unitary embedding $`A`$. $`\mathrm{}`$ Example. Not all deterministic patterns are uniformly or strongly so. To see this, choose as command sequence $`X_1^{s_2}M_2^0E_{12}N_2^0`$, with $`V=\{1,2\}`$, and $`I=O=\{1\}`$. The two branch maps are given by $`A_0=|00|`$, and $`A_1=|01|`$, so they are proportional, but distinct, and the pattern is deterministic, but not strongly so. The associated cptp-map $`T(|\psi \psi |)=\psi ,\psi |00|`$ projects any state onto $`|0`$ and does not correspond to a unitary transformation. This pattern is not uniformly deterministic either, since $`\alpha =0`$ is the only angle value for $`M_2^\alpha `$ which makes it deterministic. ## III Geometries and Flows An *open graph states* $`(G,I,O)`$ consists of an undirected graph $`G`$ together with two subsets of nodes $`I`$ and $`O`$, called inputs and outputs. We write $`V`$ for the set of nodes in $`G`$, $`I^c`$, and $`O^c`$ for the complements of $`I`$ and $`O`$ in $`V`$, $`G(i)`$ for the set of neighbors of $`i`$ in $`G`$, and $`E_G:=_{(i,j)G}E_{ij}`$ for the global entanglement operator associated to $`G`$. One may think of an open graph state as the beginning of the definition of a pattern, where one has already decided how many qubits will be used ($`V`$), how they will be entangled ($`E_G`$), and which will be inputs and which outputs ($`I`$ and $`O`$). To complete the definition of the pattern it remains to decide which angles will be used to prepare qubits in $`I^c`$ (qubits in $`I`$ are given in an arbitrary states) which angles will be used to measure qubits in $`O^c`$, and most importantly, if one is interested in determinism, which dependent corrections will be used. Conversely, any pattern has a unique underlying open graph state, obtained by forgetting preparations, measurements and corrections. For instance, the open graph state associated to the example above is the graph $`G`$ with nodes $`\{1,2\}`$, inputs and outputs $`\{1\}`$, and $`E_G=E_{12}`$. To complete the definition, one has to choose the angles of the measurement and preparations done at qubit $`2`$, and define the dependent corrections. We give a condition bearing on the geometry of open graph states, under which one can construct a set of dependent corrections such that the obtained pattern is strongly and uniformly deterministic. ###### Definition 2 An open graph state $`(G,I,O)`$ has *flow* if there exists a map $`f:O^cI^c`$ (from measured qubits to prepared qubits) and a partial order $`>`$ over $`V`$ such that for all $`iO^c`$: — (F0) $`(i,f(i))G`$ — (F1) $`f(i)>i`$ — (F2) for all neighbours of $`f(i)`$ except $`i`$ ($`kG(f(i))\{i\}`$), we also have $`k>i`$ As one can see, a flow consists of two structures: a function $`f`$ over vertices and a matching partial order over vertices. In order to obtain a deterministic pattern for an open graph state with flow, dependent corrections will be defined based on function $`f`$. The order of the execution of the commands is given by the partial order induced by the flow. The matching properties between the function $`f`$ and the partial order $`>`$ will make the obtained pattern runnable. Figure 1 shows an open graph state together with a flow, where function $`f`$ represented as arrows from $`O^c`$ (measured qubits, black circles) to $`I^c`$ (prepared qubits, non boxed nodes). The associated partial order is given by the labeled sets of vertices. The coarsest order $`>`$ such that (F1) and (F2) holds is called the *dependency order* induced by the flow, and the number of the partition sets (4 in Figure 1) is called the *depth* of the flow. In general flows may or may not exist, and are not unique either. ###### Theorem 1 Suppose the open graph state $`(G,I,O)`$ has flow $`(f,>)`$, then the pattern: $$\begin{array}{ccccccccccccccc}𝔓_{f,G,>,\stackrel{}{\alpha }}\hfill & :=\hfill & _{iO^c}^>(X_{f(i)}^{s_i}_{kG(f(i))\{i\}}Z_k^{s_i}M_i^{\alpha _i})E_GN_{I^c}^0\hfill & & & & & & & & & & & & \end{array}$$ where the product follows the dependency order $`>`$, is runnable, uniformly and strongly deterministic, and realizes the unitary embedding: $$\begin{array}{ccccccccccccccc}U_{G,I,O,\stackrel{}{\alpha }}\hfill & :=\hfill & (_{iO^c}+_{\alpha _i}|_i)E_GN_{I^c}^0\hfill & & & & & & & & & & & & \end{array}$$ Proof. The proof is based on the following equations, where $`s`$ stands for any arbitrary $`s_j`$: $`+_\alpha |_i`$ $`=`$ $`M_i^\alpha Z_i^{s_i}`$ (1) $`Z_i^sE_{ij}`$ $`=`$ $`X_j^sE_{ij}X_j^s`$ (2) $`X_i^sE_{ij}`$ $`=`$ $`E_{ij}Z_j^sX_i^s`$ (3) $`Z_i^sE_{ij}`$ $`=`$ $`E_{ij}Z_i^s`$ (4) $`X_i^sN_i^0`$ $`=`$ $`N_i^0`$ (5) Equation (1) amounts to saying that $`Z_i|\pm _\alpha _i=|_\alpha _i`$; notice also that this property uniquely defines $`Z`$. Equations (2), (3), and (4) come from the fact that $`Z`$ is in the normalizer of the Pauli group, and are easy to verify. Equation (5) is obvious. From (1) we obtain: $$\begin{array}{ccccccccccccccc}_{iO^c}+_\alpha |_iE_GN_{I^c}^0\hfill & =_{(\text{1})}\hfill & (_{iO^c}M_i^{\alpha _i}Z_i^{s_i})E_GN_{I^c}^0\hfill & & & & & & & & & & & & \end{array}$$ so the right hand side is clearly a deterministic pattern, but just as clearly it violates condition (R0), since $`Z_i^{s_i}`$ depends on a measurement which has not been done yet. At that point, entanglement comes to rescue. Write $`G(i)^c`$ for the graph obtained by removing $`G(i)`$ from $`G`$. Then we can rewrite the above pattern as follows, where boxes represent the part to which we apply the rewriting equations: $$\begin{array}{ccccccccccccccc}Z_i^{s_i}E_GN_{I^c}^0\hfill & =\hfill & & & & & & & & & & & & & \\ Z_i^{s_i}E_{G(i)}E_{G(i)^c}N_{I^c}^0\hfill & =\hfill & & & & & & & & & & & & & \\ \overline{)Z_i^{s_i}E_{if(i)}}(_{kG(f(i))\{i\}})E_{f(i)k}E_{G(f(i))^c}N_{I^c}^0\hfill & =_{(\text{2})}\hfill & & & & & & & & & & & & & \\ X_{f(i)}^{s_i}E_{if(i)}\overline{)X_{f(i)}^{s_i}(_{kG(f(i))\{i\}}E_{f(i)k}})E_{G(f(i))^c}N_{I^c}^0\hfill & =_{(\text{3})}\hfill & & & & & & & & & & & & & \\ X_{f(i)}^{s_i}\overline{)E_{G(f(i))}(_{kG(f(i))\{i\}}Z_k^{s_i})}X_{f(i)}^{s_i}E_{G(f(i))^c}N_{I^c}^0\hfill & =_{(\text{4})}\hfill & & & & & & & & & & & & & \\ X_{f(i)}^{s_i}(_{kG(f(i))\{i\}}Z_k^{s_i})E_G\overline{)X_{f(i)}^{s_i}N_{I^c}^0}\hfill & =_{(\text{5})}\hfill & & & & & & & & & & & & & \\ X_{f(i)}^{s_i}(_{kG(f(i))\{i\}}Z_k^{s_i})E_GN_{I^c}^0\hfill & & & & & & & & & & & & & & \end{array}$$ Condition (F0) is used in the third step. Finally: $$\begin{array}{ccccccccccccccc}_{iO^c}+_\alpha |_iE_GN_{I^c}^0=\hfill & & & & & & & & & & & & & & \\ (_{iO^c}X_{f(i)}^{s_i}(_{kG(f(i))\{i\}}Z_k^{s_i})M_i^{\alpha _i})E_GN_{I^c}^0\hfill & & & & & & & & & & & & & & \end{array}$$ By conditions (F1) and (F2) the obtained pattern is runnable, since the product can always be ordered according to $`>`$. Moreover, by the last equation, all branch maps are equal, and therefore the pattern is strongly deterministic. Finally, since the proof uses nowhere the particular values of the measurement angles $`\alpha _i`$, it is also uniformly so. $`\mathrm{}`$ The intuition of the proof is that Equation 2 converts an anachronical $`Z`$ correction at $`i`$, given in the term $`M_i^\alpha Z_i^{s_i}`$, into a pair of a ‘future’ $`X`$ correction, the one sent to $`f(i)`$ (so in the future, by condition (F1)) and a ‘past’ $`X`$ correction, sent to the past, until it reaches a preparation, where it is absorbed because of Equation 5. Note that the unitary embedding associated to $`𝔓_{f,G}`$ (we drop $`\stackrel{}{\alpha }`$ and $`>`$, for simplicity) does not depend on the flow. Yet, the choice of $`(f,>)`$ determines the structure of the corrections used by the pattern and the order of the execution, and has therefore an influence on its depth complexity, which is the depth of the flow. Another thing worth noticing, is that using the graph stabilizer NC00 ; graphstates at $`i`$, defined as $`K_{G(i)}:=X_i(_{jG(i)}Z_j)`$, the pattern $`𝔓_{f,G}`$ can be equivalently written as: $$\begin{array}{ccccccccccccccc}𝔓_{f,G}\hfill & =\hfill & _{iO^c}^>(M_i^{\alpha _i}Z_i^{s_i}K_{G(f(i))}^{s_i})E_GN_{I^c}^0\hfill & & & & & & & & & & & & \end{array}$$ and the above proof can be reread in terms of stabilizers. In another word, for cancelling an anachronical $`Z`$ correction at $`i`$ it is enough to apply the dependent stablizer at qubit $`f(i)`$, $`K_{G(f(i))}^{s_i}`$ and again conditions $`(F1)`$ and $`(F2)`$ guarantee that the obtained pattern is runnable. ### III.1 Pauli Measurements As we saw before, not all open graph states have flow. Figure 2 shows such an example, let $`f`$ be a candidate flow function, then the only choice for $`f(a)`$ is node $`c`$, same is true for $`f(b)`$. Now from condition (F2) node $`b`$ must be in the future of node $`a`$ and vice versa. Hence we reach a causality conflict. However, one can still obtain a deterministic pattern for the open graph state in Figure 2 by fixing the angle of the measurement of node $`b`$ to be $`\pi /2`$. To see why, recall that Condition (F1) forbids $`f(i)=i`$, yet, in the special case where qubit $`i`$ is measured with angle $`\frac{\pi }{2}`$ (Pauli $`Y`$ measurement), choosing $`f(i)`$ to be $`i`$ will work, since: $$\begin{array}{ccccccccccccccc}M_i^{\frac{\pi }{2}}X_i^s=M_i^{\frac{\pi }{2}}Z_i^s\hfill & & & & & & & & & & & & & & \end{array}$$ Hence to correct the $`Y`$ measurement at qubit $`i`$ one can apply the dependent stabilizer, $`(Z_i(_{jG(i)}Z_j))^{s_i}`$, at the same qubit $`i`$ instead of a neighboring qubit, Figure 3. However the obtained pattern is deterministic only if qubit $`b`$ is measured with angle $`\frac{\pi }{2}`$, and is therefore not uniformly deterministic. Note that in the above example we fixed $`f(b)=b`$ but condition $`F(3)`$ still need to be verified. And this is indeed the case since in the given partial order the qubit $`c`$ which is neighbour of qubit $`b`$ is in the next level. To make this point clear consider the open graph state in Figure 4. The only choice for $`f(a)`$ is $`b`$ and hence $`a<b`$ but then letting $`f(b)=b`$ will violate the $`F(3)`$ condition. Therefore the only solution is to consider $`Y`$ measurement as an arbitrary measurement then we obtain a flow, Figure 5. Another special case is when qubit $`f(i)`$ is measured with angle $`0`$ (Pauli $`X`$ measurement). Again the requirement that $`f(i)>i`$ can be dropped because: $$\begin{array}{ccccccccccccccc}M_i^0X_i^s=M_i^0\hfill & & & & & & & & & & & & & & \end{array}$$ Therefore in the flow construction where the neighboring qubit $`f(i)`$ receives $`X_{f(i)}^{s_i}`$, if it is measured with angle $`0`$ this correction can be ignored. The special cases of Pauli measurements can be related to the fact that Pauli measurements transform one graph state to another one graphstates . Hence one can observe that for open graph states without flow, there might exists a set of Pauli measurements that transform it to one with flow. ### III.2 Circuit Decomposition Flow also provides a decomposition into simple building blocks, called *star patterns*, from which one can derive a corresponding circuit implementation of the pattern. Define the star pattern $`𝔖(n,\alpha )`$ as: $$\begin{array}{ccccccccccccccc}& X_2^{s_1}M_1^\alpha E_{12}E_{13}\mathrm{}E_{1n}\hfill & & & & & & & & & & & & & \end{array}$$ where 1 is the only input and $`2,\mathrm{},n`$ are the outputs, for $`n2`$. The underlying graph has a simple flow function with $`f(1)=2`$ and a two level partial order (see Figure 6). It is easy to verify that the Star pattern implements the unitary given by the circuit in Figure 7. Every pattern such that the underlying open graph state has flow can be decomposed into star patterns. The construction starts by picking a qubit in the first level of the partial order, exhausts all qubits in the first level before going to the next level. Each time a qubit $`i`$ is picked the associated star pattern $`S_i`$ is taken to have $`i`$ as input, and all its remaining current neighbours as outputs. Then we remove this qubit from the graph and carry on the construction till we reach to the final level of partial order. The final deterministic pattern is the sequential and tensor composition of the obtained star patterns with the final $`Z`$ between the output qubits: $$\begin{array}{ccccccccccccccc}𝔓=_{m,nO}E_{mn}_{iO^c}^>S_i\hfill & & & & & & & & & & & & & & \end{array}$$ Now each Start pattern can be replaced by its corresponding circuit to give a circuit decomposition for the pattern $`𝔓`$. In the obtained circuit each wire represents either an input qubit or an auxillary one prepared in $`|+`$ state, where the case is determined during the above construction. This construction can be easily formalized. ## IV Algebraic Structure As yet, Theorem 1 is only valid when preparations are all of the form $`N_i^0`$ since Equation (5) in the proof is valid only for such preparations. Define $`X_i^\alpha =Z_i^\alpha X_iZ_i^\alpha `$, with $`Z_i^\alpha `$ the phase operator with angle $`\alpha `$ applied at $`i`$. One has $`Z_i=Z_i^\pi `$. To handle general phase preparations, one only needs the analog of equations (2), (3) and (5): $$\begin{array}{ccccccccccccccc}Z_i^sE_{ij}\hfill & =\hfill & (X_i^\alpha ){}_{}{}^{s}E_{ij}^{}(X_i^\alpha )^s\hfill & & & & & & & & & & & & \\ (X_i^\alpha )^sE_{ij}\hfill & =\hfill & E_{ij}Z_j^s(X_i^\alpha )^s\hfill & & & & & & & & & & & & \\ (X_j^\alpha )^sN_i^\alpha \hfill & =\hfill & N_i^\alpha \hfill & & & & & & & & & & & & \end{array}$$ and now Theorem 1 works as before. Note that we had to extend the set of corrections to include $`X_i^\alpha `$. This extension will prove natural below, when we deal with adjunction. Say an open graph state $`(G,I,O)`$ has *bi-flow*, if both $`(G,I,O)`$ and its dual state $`(G,O,I)`$ have flow. Say a pattern has flow (bi-flow) if its underlying open graph state does. The class of patterns with flows (bi-flows) is closed under composition and tensorization. It is also universal, in the sense that all unitaries can be realised within this class. This follows from the existence of a set of generating patterns having bi-flow generator04 . Figure 8 shows the open graph state corresponding to a pattern realizing $`U`$ (controlled-$`U`$), for $`U`$ an arbitrary 1-qubit unitary generator04 . Patterns with bi-flows realize unitary operators. Indeed, by (F2), a flow $`(f,>)`$ is one-to-one. Therefore the orbits $`f^n(i)`$ for $`iI`$ define an injection from $`I`$ into $`O`$. In the case of a bi-flow, $`I`$ and $`O`$ are therefore in bijection, and since one knows already that patterns with flows realize unitary embeddings, it follows that patterns with bi-flow implement unitaries. Interestingly, one can define directly the adjoint of a pattern in the subcategory of patterns with bi-flows. Specifically, given $`(f,>)`$ a flow for $`(G,I,O)`$, and angles $`\{\alpha _i;iI^c\}`$ for preparations, and $`\{\beta _j;jO^c\}`$ for measurements, we write $`𝔓_{f,G,\stackrel{}{\alpha },\stackrel{}{\beta }}`$ for the pattern obtained as in the extension to general preparations of Theorem 1. Suppose a reverse flow $`(g,>)`$ is given on $`(G,O,I)`$, one can define: $$\begin{array}{ccccccccccccccc}𝔓_{f,G,\stackrel{}{\alpha },\stackrel{}{\beta }}^{}\hfill & :=\hfill & 𝔓_{g,G,\stackrel{}{\beta },\stackrel{}{\alpha }}\hfill & & & & & & & & & & & & \end{array}$$ There are two things to note here: first, for this definition to make sense, one needs to have general preparations as we described above; second, this adjunction operation depends on the choice of a reverse flow $`(g,>)`$. It is easy to see that $`𝔓_{f,G,\alpha ,\beta }^{}`$ and $`𝔓_{f,G,\beta ,\alpha }`$ realize adjoint unitaries. An example is the pattern $`:=X_2^{s_1}M_1^0E_{12}N_2^0`$ with $`I=\{1\}`$ and $`O=\{2\}`$. It has a unique bi-flow, and is self-adjoint in the sense that $`^{}=`$, therefore it must realize a self-adjoint operator, and indeed it realizes the Hadamard transformation. ## V Conclusion Whereas the one-way model had been mostly thought of in relation with the traditional circuit model, we have proposed here a flow condition, which is clearly divorced from the circuit model, and guarantees the existence of a set of Pauli corrections obtaining a (strongly and uniformly) deterministic behavior. In essence, while dealing with patterns with flow, one can wholly forget about corrections, and think of measurements as being simply projections. This in turn may help in revealing the new perspective on quantum computing which is implicit in measurement based models. Following this work, a polynomiual time algorithm for finding flow was proposed in Beaudrap06 which then extended to an algorithmic method for circuit design for unitaries thoroughly based on the one-way model BDK06 . Furthermore one can see that given an open graph state as a resource for computation, flow condition characterizes the set of all unitaries implementable on that given state. If one is ready to lose uniform determinism, this condition can be somewhat extended when dealing with Pauli measurements. It may be however that strong and uniform determinism is an interesting property, when it comes to fault-tolerant computing in the one-way model. Another point worth making is that the notion of flow gives a better understanding of why $`X^\alpha `$, $`Z`$ corrections and $`N^\alpha `$ preparations are needed. From the point of view of our determinism theorem (Theorem 1), they represent a natural and universal way to control the non deterministic evolutions induced by 1-qubit $`XY`$ measurements on a graph state. Finally, although the obtained class of patterns with flow is universal, it remains to be seen whether this condition is also necessary for determinism. One also need to extend the flow condition to deal with $`XZ`$ and $`YZ`$ plane measurements, which are the topics of our future work. ## Acknowledgments The authors wish to thank Niel de Beaudrap, Anne Broadbent, Ignacio Cirac, Paul Dumais, Damian Markham, Keiji Matsumoto for useful comments and discussions. EK was partially supported by the ARDA, MITACS, ORDCF, and CFI projects during her stay at Institute for Quantum Computing at University of Waterloo.
warning/0506/astro-ph0506482.html
ar5iv
text
# Soft coincidence in late acceleration ## I Introduction Nowadays it is widely accepted that the present stage of cosmic expansion is accelerated reviews ; adam albeit there are rather divergent proposals about the mechanism behind this acceleration. A cosmological model of present acceleration should not only fit the high redshift supernovae data, the cosmic microwave background anisotropy spectrum and safely pass other tests, it must solve the coincidence problem as well, namely “why the Universe is accelerating just now?”, or in the realm of Einstein gravity “why are the densities of matter and dark energy of precisely the same order today?” coincidence -note that these two energies scale differently with redshift. While it might happen that this coincidence is just a “coincidence” -and as such no explanation is to be found- we believe models that fail to account for this cannot be regarded as satisfactory. In a class of models designed to solve this problem the dark energy density “tracks” the matter energy density for most of the history of the Universe, and overcomes it only recently (see, e.g., Ref. trackers ). However, these models suffer the drawback of fine-tuning the initial conditions whereupon they are not, after all, much better than the conventional “concordance” model which rests on a mixture of matter and a fine-tuned cosmological constant sabino . There is an especially successful subset of models based on an interaction between dark energy and cold matter (i.e., dust) such that the ratio $`r`$ of the corresponding energy densities tends to a constant of order unity at late times plb ; luca ; scale ; interacting ; dw ; somasri ; olivares thus solving the problem. However, the current observational information does not necessarily imply that $`r`$ ought to be strictly constant today. For the coincidence problem to be addressed a softer condition may suffice, namely that at present $`r`$ should be a slowly varying function of the scale factor with $`r(a=a_0)3/7`$, the currently observed ratio. By slowly varying we mean that the current rate of variation of $`r(a)`$ should be no much larger than $`H_0`$, where $`H\dot{a}/a`$ denotes the Hubble factor of the Friedamnn-Lemaitre-Robertson-Walker (FLRW) metric, and a zero subscript means present time. It should be noted that because the nature of dark matter and dark energy is largely unknown an interaction between both cannot be excluded a priori. In fact, the possibility has been suggested from a variety of angles angles . To avoid a possible conflict with observational constraints on long-range forces only we consider that the baryon component of the matter does not participate in the the interaction and, further, to simplify the analysis -i.e., in order not to have an uncoupled component- we exclude the baryons altogether. While this might be seen as a radical step it should be taken into account that our study restricts itself to times near the present time and these are characterized, among other things, by a low value of the baryon energy density ($`5`$% or less of the total energy budget, approximately six times lower than the dark matter contribution and fourteen below the dark energy component spergel ) whereby it should not significantly affect our results. This is in keeping with the findings of Majerotto et al. majerotto . For interacting models encompassing most the Universe history in which the baryons enter the dynamical equations as a non-interacting component, see Refs. majerotto and luca . The target of this paper is to present two models of late acceleration that fulfill “soft coincidence”, namely: $`(i)`$ when the dark energy is a quintessence scalar field, and $`(ii)`$ when the dark energy is a tachyon field. The latter was introduced by Sen sen and soon afterwards it became a candidate for driving inflation as well as late acceleration -see e.g., Ref. gary . The outline of this paper is as follows: Section II considers the quintessence model with a constant equation of state parameter. There it is assumed that the quintessence field slowly decays into dark matter with the equation of state of dust. Section III considers the tachyon field and again assumes a slowly decay into dust. This time, however, the equation of state parameter is allowed to vary. Finally, section IV summarizes our findings. ## II The quintessence interacting model We consider a two-component system, namely, cold dark matter, described by an energy density $`\rho _m`$, and a quintessence scalar field $`\varphi `$ with energy density and pressure defined by $$\rho _\varphi =\frac{1}{2}\dot{\varphi }^2+V(\varphi ),\text{and}p_\varphi =\frac{1}{2}\dot{\varphi }^2V(\varphi ),$$ (1) respectively, in a spatially flat FLRW universe. The over dot indicates derivative with respect to the cosmic time and $`V(\varphi )`$ is the quintessence scalar potential. We assume that these two components do not evolve separately but interact through a source (loss) term (say, $`Q`$) that enters the energy balances $$\dot{\rho }_m+3H\rho _m=Q$$ (2) and $$\dot{\rho }_\varphi +3H(\rho _\varphi +p_\varphi )=Q,$$ (3) where in view of (1) last equation is equivalent to $$\dot{\varphi }\left[\ddot{\varphi }+3H\dot{\varphi }+V^{}\right]=Q.$$ (4) In the following we constrain the interaction $`Q`$ by demanding that the solution to Eqs. (2) and (3) be compatible with a variable ratio between the energy densities $`r(x)\rho _m/\rho _\varphi `$, where $`x=a/a_0`$ is the normalized scale factor, and that around the present time $`r(x)`$ is a smooth, nearly constant function with $`r(x=1)=r_0`$ being of order one. We also assume that the quintessence component obeys a barotropic equation of state, it is to say $`p_\varphi =w_\varphi \rho _\varphi `$ with $`w_\varphi `$ a negative constant (a distinguishing feature of dark energy fields -quintessence fields or whatever- is a high negative pressure). In virtue of these relations the set of dynamical equations reduces to a single equation $$\dot{\rho }_\varphi +3H\left(1+\frac{w_\varphi +\frac{\dot{r}}{3H}}{1+r}\right)\rho _\varphi =0,$$ (5) whose solution is $$\rho _\varphi (x)=\rho _\varphi ^{(0)}e^{3I(x)},$$ (6) with $$I(x)=_1^xF(x^{})\frac{dx^{}}{x^{}},\text{where}F(x)=1+\frac{w_\varphi +\frac{1}{3}xr^{}(x)}{1+r(x)},$$ (7) and the prime means derivation with respect to $`x`$. On the other hand, by combining Friedmann’s equation $$3H^2=\kappa (\rho _m+\rho _\varphi )(\kappa 8\pi G)$$ (8) with Eq. (6) we get $$H(x)=H_0\sqrt{\frac{1+r(x)}{1+r_0}}e^{\frac{3}{2}I(x)},$$ (9) where $`H_0=\sqrt{\rho _\varphi ^{(0)}\kappa /3}`$ denotes the current value of the Hubble factor. From this, it follows that $$\frac{H_0}{\sqrt{1+r_0}}\left[t(x)t_0\right]=_1^x\frac{e^{\frac{3}{2}I(x^{})}}{\sqrt{1+r(x^{})}}\frac{dx^{}}{x^{}}.$$ (10) If this integral could be solved analytically, we would obtain the scale factor in terms of the cosmological time. Equations (3) and (5) alongside (6), (7), and (9) imply $$Q(x)=\frac{3H_0\rho _\varphi ^{(0)}}{\sqrt{1+r_0}}\frac{1}{\sqrt{1+r(x)}}\left[\frac{xr^{}(x)}{3}w_\varphi r(x)\right]e^{{\scriptscriptstyle \frac{9}{2}}I\left(x\right)}.$$ (11) Note that $`Q(x)`$ is a positive-semidefinite function, as it should. A negative $`Q(x)`$ would imply a transfer of energy from the matter to the scalar field which migth violate the second law of thermodynamics. While in view of the unknown nature of dark matter and dark energy we cannot say for certain that these components fulfill the aforesaid law, in the absence of any evidence against it, the most natural thing is to assume that they obey it. From the definitions (1) and the equation of state $`p_\varphi =w_\varphi \rho _\varphi `$ the quintessence field and its potential are given by $$\varphi (x)=\varphi _0+\sqrt{\frac{3(1+w_\varphi )}{\kappa }}_1^x\sqrt{\frac{1}{1+r(x^{})}}\frac{dx^{}}{x^{}},$$ (12) and $$V(x)=V_0e^{3I(x)},$$ (13) respectively. Here $`\varphi _0`$ is an integration constant, and $`V_0=\frac{1}{2}\left(1w_\varphi \right)\rho _\varphi ^{(0)}`$. As said before, we apply the above formalism to the case in which the variable $`x`$ is not far from unity whence the ratio $`r(x)`$ can be approximated by $$r(x)r_0+\epsilon _0(1x),$$ (14) where $`r_0`$ is the present value of the ratio between the energy densities $`\rho _m`$ and $`\rho _\varphi `$, and $`\epsilon _0`$ is a small positive-definite constant. We do not consider negative values for $`\epsilon _0`$ since it would imply that $`r(x)`$ was increasing in the recent past and therefore that it oscillates. While we are unaware of any definitive argument against this possibility it looks certain that only contrived models may lead to this behavior. Further, oscillations in $`r(x)`$ may seriously jeopardize the well tested picture of structure formation springel . (Note in passing that the choice (14) implies that $`\dot{r}(x)H_0`$ for $`x𝒪(1)`$). It follows that $$F(x)=1+\frac{\alpha _1\alpha _2x}{\alpha _3\alpha _4x},$$ (15) where the $`\alpha _i`$ are constants given by $$\alpha _1=w_\varphi ,\alpha _2=\frac{1}{3}\epsilon _0,\alpha _3=1+r_0+\epsilon _0,\alpha _4=\epsilon _0,$$ (16) respectively. With this, we obtain $$\rho _m(x)=\rho _\varphi ^{(0)}\left(\frac{\alpha _3\alpha _4}{\alpha _3\alpha _4x}\right)(\alpha _31\alpha _4x)x^{3\left(\frac{\alpha _1+\alpha _3}{\alpha _3}\right)}.$$ (17) and $$\rho _\varphi (x)=\rho _\varphi ^{(0)}\left(\frac{\alpha _3\alpha _4}{\alpha _3\alpha _4x}\right)x^{3\left(\frac{\alpha _1+\alpha _3}{\alpha _3}\right)},$$ (18) The Hubble parameter can be written as $$H(x)=H_0x^{\frac{3}{2}\left(\frac{\alpha _1+\alpha _3}{\alpha _3}\right)},$$ (19) where $`H_0`$ was given above. From this expression the scale factor is shown to follow a power-law dependence on time $$a(t)=a_0\left(\frac{t}{t_0}\right)^{\frac{2\alpha _3}{3(\alpha _1+\alpha _3)}}.$$ (20) For the Universe to accelerate, the constraint $`\frac{2\alpha _3}{3\left(\alpha _1+\alpha _3\right)}>1`$ must be fulfilled, i.e., $$1+r_0+\epsilon _0<3w_\varphi .$$ (21) This, alongside the condition that the energy densities decrease with expansion implies $$w_\varphi <1+r_0+\epsilon _0<3w_\varphi .$$ (22) Fig. 1 shows a good fit to the “gold” set of supernovae data points of Riess et al. adam . The likelihood contours are depicted in Fig. 2. The mean values of the free parameters are: $`\mathrm{\Omega }_\varphi =0.78538`$, $`w_\varphi =0.757284`$, $`\epsilon _0=0.0777764`$. Notice that $`\mathrm{\Omega }_\varphi `$ is above the concordance $`\mathrm{\Lambda }`$CDM value of $`0.73`$ and that $`w_\varphi `$ is significantly larger than the value found in non-interacting models. However, we have not considered phantom fields (scalar fields as given by Eq. (1) do not encompass phantom behavior), otherwise a shift of $`w_\varphi `$ toward more negative values should be expected. Notice (top panels) that the parameter $`\epsilon _0`$ is rather degenerate. Under restriction (21) the interaction term reads $$Q(x)=3H_0\rho _\varphi ^{(0)}\left[w_\varphi +\frac{\alpha _4x}{3[(\alpha _31\alpha _4x)]}\right](\alpha _3\alpha _4)\frac{\left(\alpha _31\alpha _4x\right)}{\left(\alpha _3\alpha _4x\right)^2}x^{{\scriptscriptstyle \frac{9}{2}}\left({\scriptscriptstyle \frac{\alpha _1+\alpha _3}{\alpha _3}}\right)},$$ (23) see Fig. 3. In its turn, the scalar field $`\varphi `$ and the scalar potential $`V(x)`$ are given by $$\varphi (x)=\varphi _0\left[\frac{\mathrm{tanh}^1\left(\sqrt{1\frac{\alpha _4}{\alpha _3}x}\right)}{\mathrm{tanh}^1\left(\sqrt{1\frac{\alpha _4}{\alpha _3}}\right)}\right],$$ (24) and $$V(x)=V_0\left(\frac{\alpha _3\alpha _4}{\alpha _3\alpha _4x}\right)x^{3\left(\frac{\alpha _1+\alpha _3}{\alpha _3}\right)},$$ (25) respectively. Here $`\varphi _0=\sqrt{\frac{12}{\kappa }\left(\frac{1+w_\varphi }{\alpha _3}\right)}\mathrm{tanh}^1\left[\sqrt{1\frac{\alpha _4}{\alpha _3}}\right]`$, and $`V_0=\frac{1}{2}(1w_\varphi )\rho _\varphi ^{(0)}`$, see Fig. 4. Equations (24) and (25) lead to $$V(\varphi )=\frac{V_0\left(1\frac{\alpha _4}{\alpha _3}\right)}{\mathrm{tanh}^2\left(\sqrt{\frac{\kappa \alpha _3}{12\left(1+w_\varphi \right)}}\varphi \right)}\left[\sqrt{\frac{\alpha _3}{\alpha _4}}\text{sech}\left(\sqrt{\frac{\kappa \alpha _3}{12\left(1+w_\varphi \right)}}\varphi \right)\right]^{6\left(\frac{\alpha _1+\alpha _3}{\alpha _3}\right)}.$$ (26) Figures 4 and 5 taken together show that the potential decreases with the Universe expansion. While we do not know whether there is any field theory backing this potential it is intriguing to see that around $`\varphi =0`$ it behaves as $$V(\varphi )C_1\varphi ^2+C_2+C_3\varphi ^2+C_4\varphi ^4+\mathrm{}$$ (27) where the $`C_i`$’s are constants. The first term is used in quitessence models -see e.g., Ref. potentials , whereas the third and fourth terms of the expansion are well–known potentials in inflation theory (chaotic potentials) Linde ; the second term plays the role of a cosmological constant. This leads us to surmise that, in reality, $`V(\varphi )`$ might be considered an effective potential resulting from the combination of a number of fields. ## III The tachyon interacting model The tachyon field naturally emerges as a straightforward generalization of the Lagrangian of the relativistic particle much in the same way as the scalar field $`\varphi `$ arises from generalizing the Lagrangian of the non–relativistic particle bagla . Recently, the realization of its potentiality as dark matter shiu and dark energy awakened has awakened the interest in it. We begin by succinctly recalling the basic equations of the tachyon field to be used below where its interaction with cold dark matter (dust) will be considered. The stress–energy tensor of the tachyon field $$T_{ab}^{(\phi )}=\frac{V(\phi )}{\sqrt{1+\phi ^{,c}\phi _{,c}}}\left[g_{ab}\left(1+\phi ^{,c}\phi _{,c}\right)+\phi _{,a}\phi _{,b}\right],$$ (28) admits to be written in the perfect fluid form $$T_{ab}^{(\phi )}=\rho _\phi u_au_b+p_\phi \left(g_{ab}+u_au_b\right),$$ (29) where the energy density and pressure are given by $$\rho _\phi =\frac{V(\phi )}{\sqrt{1\dot{\phi }^2}},\text{and}p_\phi =V(\varphi )\sqrt{1\dot{\phi }^2},$$ (30) respectively, with $$\dot{\phi }\phi _{,a}u^a=\sqrt{g^{ab}\phi _{,a}\phi _{,b}},\text{and}u_a=\frac{\phi _{,a}}{\dot{\phi }},\text{where}u^au_a=1.$$ (31) In the absence of interactions other than gravity the evolution of the energy density is governed by $`\dot{\rho }_\phi =3H\dot{\phi }^2\rho _\phi `$, therefore when $`\dot{\phi }^2<1`$ it decays at a lower rate than that for dust. It approaches the behavior of dust for $`\dot{\phi }^21`$, thereby in this limit the tachyon field behaves dynamically as pressureless matter does. Consequently, we shall assume $`\dot{\phi }^2<1`$ since for $`\dot{\phi }^2=1`$ both components obey the same equation of state for dust. For an interacting mixture of a tachyon field and cold dark matter, with energy density $`\rho _m`$ and negligible pressure, the interaction term, $`Q`$, between these two components is described by the following balance equations $$\dot{\rho }_m+3H(\rho _m+\mathrm{\Pi }_m)=Q,$$ (32) $$\dot{\rho }_\phi +3H\dot{\phi }^2\rho _\phi =Q.$$ (33) The $`\mathrm{\Pi }_m`$ term, on the left hand side of Eq. (32), accounts for the fact that the matter component may be endowed with a viscous pressure or perhaps it is slowly decaying into dark matter and/or radiation decay . In either case one can model this term as $`\mathrm{\Pi }_m=\alpha \rho _mH`$ with $`\alpha `$ a small negative constant since $`\mathrm{\Pi }_m`$ is a small correction to the matter pressure -see winfried and references therein. As before, we consider the ratio between the densities of matter and tachyonic energy a function $`r(x)`$ of the normalized scale factor (to be specified later), and again, we must have $`Q(x)>0`$ -its expression is to be found below. Then, equations (32) and (33) combine to $$\dot{\rho }_\phi +3H\left[1+\frac{w_\phi +\alpha r(x)+\frac{r^{}(x)x}{3}}{1+r(x)}\right]\rho _\phi =0(xa/a_0).$$ (34) The latter can be solved to $$\rho _\phi (x)=\rho _\phi ^{(0)}e^{3\stackrel{~}{I}(x)},\text{where}\stackrel{~}{I}(x)=_1^x\stackrel{~}{F}(x^{})\frac{dx^{}}{x^{}},$$ (35) with $$\stackrel{~}{F}(x)=1+\frac{w_\phi (x)+\alpha r(x)+\frac{1}{3}r^{}(x)x}{1+r(x)}.$$ (36) The interaction term takes the form $$Q(x)=3\rho _\phi ^{(0)}H(x)\left(\frac{r(x)}{1+r(x)}\right)\left[\alpha w_\phi (x)+\frac{xr^{}(x)}{3r(x)}\right]e^{3\stackrel{~}{I}(x)}.$$ (37) and the tachyonic scalar field and its potential obey $$\phi (x)=\phi _0+\frac{\sqrt{1+r_0}}{H_0}_1^x\sqrt{\frac{1+w_\phi (x^{})}{1+r(x^{})}}e^{\frac{3}{2}\stackrel{~}{I}(x^{})}\frac{dx^{}}{x^{}},$$ (38) and $$V(x)=V_0\sqrt{w_\phi (x)}e^{3\stackrel{~}{I}(x)},$$ (39) respectively. Up to now we have left the ratio function $`r(x)`$ free. As before, we specify it for $`x`$ values around unity as $$r(x)r_0+\epsilon _0(1x),$$ (40) where $`r_0=(\rho _m/\rho _\phi )_0`$, and $`\epsilon _0`$ is once again a small positive-definite constant. Likewise, we assume that the equation of state parameter $`w_\phi `$ is given by $$w_\phi (x)=w_0+w_1(1x),$$ (41) where $`w_0`$ and $`w_1`$ are constants, the first one denotes the current value of the $`w_\phi (x)`$ function, and the second one is minus its first derivative, which is expected to be small. Thus, $$\stackrel{~}{F}(x)=1+\frac{a_1b_1x}{a_2b_2x},$$ (42) where the constants $`a_i`$ and $`b_i`$ stand for $$a_1=\alpha r_0+w_0+w_1+\epsilon _0,a_2=1+r_0+\epsilon _0$$ (43) and $$b_1=w_1+(\alpha +\frac{1}{3})\epsilon _0,b_2=\epsilon _0,$$ (44) respectively. It follows that $$\rho _m(x)=\rho _\phi ^{(0)}(1+r_0)^{3\beta _1}x^{3\beta _2}[r_0+\epsilon _0(1x)][1+r_0+\epsilon _0(1x)]^{3\beta _1},$$ (45) as well as $$\rho _\phi (x)=\rho _\phi ^{(0)}(1+r_0)^{3\beta _1}x^{3\beta _2}[1+r_0+\epsilon _0(1x)]^{3\beta _1},$$ (46) with $$\beta _1=\frac{b_1}{b_2}\frac{a_1}{a_2}=\frac{w_1}{\epsilon _0}+\left(\alpha +\frac{1}{3}\right)\left[\frac{\alpha r_0+w_1+w_0+\epsilon _0}{1+r_0+\epsilon _0}\right],$$ and $$\beta _2=1+\frac{a_1}{a_2}=\frac{1+w_1+2\epsilon _0+r_0(1+\alpha )}{1+r_0+\epsilon _0}.$$ The Hubble function $$H(x)=H_0(1+r_0)^{\frac{3\beta _11}{2}}x^{\frac{3\beta _2}{2}}[1+r_0+\epsilon _0(1x)]^{\frac{(3\beta _1+1)}{2}},$$ (47) follows from the Friedmann’s equation. Although last expression is comparatively simple, the scale factor derived from it is not $$\frac{3\beta _2}{2}(1+r_0)^{\frac{3\beta _11}{2}}H_0(tt_0)=x^{\frac{3\beta _2}{2}}(1+r_0+\epsilon _0(1x))^{\frac{3\beta _11}{2}}$$ $$\times _2F_1([\frac{3\beta _2}{2},\frac{13\beta _1}{2}],\left[1+\frac{3\beta _2}{2}\right];\frac{\epsilon _0x}{1+r_0+\epsilon _0})C_1,$$ (48) where $`{}_{2}{}^{}F_{1}^{}`$ is the hypergeometric function hyperg and $$C_1=\left[(1+r_0)\left(1\frac{\epsilon _0}{1+r_0+\epsilon _0}\right)^1\right]_2^{\frac{3\beta _11}{2}}F_1(x=1).$$ Fig. 6 portrays the evolution of the scale factor in terms of the cosmological time as well as the deceleration factor $`q\ddot{a}/(aH^2)`$ versus the redshift for two selected values of the parameters. As Fig. 7 shows the model fits the supernova data points not less well than the concordance $`\mathrm{\Lambda }`$CDM model does. The likelihood contours, Figs. 8 and 9, were calculated with the method of Markov’s chains. We used the prior $`\mathrm{\Omega }_m+\mathrm{\Omega }_\phi =1`$ and that the parameters $`w_0`$ and $`w_1`$ are restricted by the condition that the value of the right hand side of Eq. (41) must lay in the interval $`[1,1/3)`$. The mean values of the parameters are: $`\mathrm{\Omega }_\phi =0.246`$, $`w_0=0.773`$, $`w_1=0.22`$, $`\epsilon _0=0.0087`$, $`\alpha =0.76`$. Here, $`\epsilon _0`$ is not so weakly constrained by the supernovae data as in the quintessence model. The present model predicts a mild evolution of the equation of state parameter with redshift. This is slightly at variance with the findings of Jassal et al. Jassal , but agrees with the model independent analysis of Alam et al. alam . The interaction term is given by $$Q(x)=Q_0\frac{(1+r_0)^{\frac{9\beta _11}{2}}}{[3r_0(\alpha w_0)\epsilon _0]}\left\{3[\alpha w_0w_1(1x)][r_0+\epsilon _0(1x)]\epsilon _0x\right\}$$ $$\times x^{\frac{9\beta _2}{2}}\left\{1+r_0+\epsilon _0(1x)\right\}^{\frac{19\beta _1}{2}},$$ (49) with $`Q_0=\frac{1}{2}\rho _\phi ^{(0)}H_0[3r_0(\alpha w_0)\epsilon _0]`$. Likewise, the tachyon field and the potential are found to be $$\phi (x)=\phi _0+\frac{(1+r_0)^{\frac{13\beta _1}{2}}}{H_0}_1^xx^{\frac{3\beta _2}{2}1}\sqrt{1+w_0+w_1(1x^{})}(1+r_0+\epsilon _0(1x^{}))^{\frac{3\beta _11}{2}}𝑑x^{}$$ (50) and $$V(x)=\rho _\phi ^{(0)}(1+r_0)^{3\beta _1}\sqrt{w_0w_1(1x)}x^{3\beta _2}[1+r_0+\epsilon _0(1x)]^{3\beta _1},$$ (51) respectively -see Fig. 10. ## IV Concluding remarks We have studied two models of late acceleration by assuming $`(i)`$ that dark energy and non-relativistic dark matter do not conserve separately but the former decays into the latter as the Universe expands, and $`(ii)`$ that the present ratio of the dark matter density to dark energy density varies slowly with time, i.e., $`\dot{r}_0H_0`$. This second assumption is key to determine the interaction $`Q`$ between both components. In the quintessence model (section II) we have considered the equation of state parameter constant while in the tachyon field model (section III) we have allowed it to vary slightly. Actually, there is no compelling reason to impose that this parameter should be a constant. However, Jassal et al. Jassal have pointed out that the WMAP data spergel imply that in any case it cannot vary much. By contrast, Alam et al. using the sample of “gold” supernovae of Riess et al. adam find a clear evolution of $`w`$ in the redshift interval $`0z1`$; however when strong priors on $`\mathrm{\Omega }_{m0}`$ and $`H_0`$ are imposed this result weakens. Nevertheless, the analysis of these two papers assume that the two main components (matter and dark energy) do not interact with each other except gravitationally. The parameter $`w_0`$ presents degeneration in both models, therefore we must wait for further and more accurate SNIa data, perhaps from the future SNAP satellite, or to resort to complementary observations of the CMB. In both cases (quintessence and tachyon), we have found analytical expressions for the relevant quantities (i.e., the scale factor, the field and the potential) and the solutions are seen to successfully pass the magnitude-redshift supernovae test -see Figs. 1 and 7. Nevertheless, it is apparent that the the tachyon model favors rather high values of the matter density parameter (see bottom right panel of Fig. 8) which is at variance with a variety of measurements of matter abundance at cosmic scales peebles which, taken as a whole, hint that $`\mathrm{\Omega }_m`$ should not exceed $`0.45`$. In consequence, the quintessence model appears favored over the tachyon model. Our work may serve to build more sophisticated models aimed to simultaneously account for the present acceleration and the coincidence problem. Previous studies of interacting dark energy aimed to solve the coincidence problem by demanding that the ratio $`r`$ be strictly constant at late times needed to prove the stability of $`r`$ at such times. This was achieved by showing that the models satisfied an attractor condition that involved the equation of state parameter of matter and dark energy as well as the Hubble factor and its temporal derivative plb ; interacting ; dw . Since in the case at hand the coincidence problem is solved with a (slowly) varying ratio $`r`$ no stability proof is necessary at all and no attractor condition is needed. Our analysis was confined to times not far from the present (i.e., for $`x𝒪(1)`$. To recover the evolution of the Universe at earlier times ($`x1`$), when the matter density dominated and produced via gravitational instability the cosmic structures we observe today, we must generalize our study along the lines of Refs. interacting and dw and include the baryon component in the dynamic equations as an uncoupled fluid. We restricted ourselves to scenarios satisfying $`w>1`$. Scenarios with $`w<1`$ (the so-called “phantom” energy models) violate the dominant energy condition though, nevertheless, they are observationally favored rather than excluded alessandro and exhibit interesting features caldwell that might call for “new physics”. We defer the study of phantom models presenting soft coincidence to a future publication. ###### Acknowledgements. Thanks are due to Germán Olivares for his computational assistance. DP is grateful to the “Instituto de Física de la PUCV” for financial support and kind hospitality. SdC was supported from Comisión Nacional de Ciencias y Tecnología (Chile) through FONDECYT grants N<sup>0</sup>s 1030469, 1010485 and 1040624 as well as by PUCV under grant 123.764/2004. This work was partially supported by the old Spanish Ministry of Science and Technology under grant BFM–2003–06033, and the “Direcció General de Recerca de Catalunya” under grant 2001 SGR–00186.
warning/0506/quant-ph0506170.html
ar5iv
text
# Bounds on Multipartite Entangled Orthogonal State Discrimination Using Local Operations and Classical Communication thanks: Authors are listed alphabetically. Corresponding authors are Markham (markham@phys.s.u-tokyo.ac.jp), Virmani (s.virmani@imperial.ac.uk) and Owari (owari@eve.phys.s.u-tokyo.ac.jp). ## Abstract We show that entanglement guarantees difficulty in the discrimination of orthogonal multipartite states locally. The number of pure states that can be discriminated by local operations and classical communication is bounded by the total dimension over the average entanglement. A similar, general condition is also shown for pure and mixed states. These results offer a rare operational interpretation for three abstractly defined distance like measures of multipartite entanglement. The problem of defining and understanding multiparty entanglement is a major open question in the field of quantum information. As entanglement theory becomes more useful in other areas of many body physics, multiparty entanglement becomes increasingly relevant to general physics, too. Hence, understanding the meaning of entanglement has become and interesting and important question. In the bipartite case, entanglement is fairly well understood Horodecki01 . There are many entanglement measures defined both operationally (in terms of the usefulness of states for quantum information tasks) and abstractly (such that they obey certain axioms and may be called entanglement monotones). One of the most celebrated results in bipartite entanglement theory is that for pure states essentially all measures coincide and have clear operational relevance. For more than two parties however, the operational approach quickly becomes very difficult. There are no clear “units of usefulness” and we have the possibility of inequivalent types of entanglement Dur00 . Some abstract measures do persist by their simplicity. In particular those measures that define “proximity” to the set of separable states Vidal99 ; Vedral98 ; Wei03 have natural multiparty analogues. However, due to their abstract definition, their operational meaning is not clear and remains an open question. In this Letter, we consider the connection between distance-like entanglement measures and the task of local operations and classical communication (LOCC) state discrimination with this question in mind. This task illustrates the restriction of only having local access to a system, fundamental to the use of entanglement in quantum information (and notions of locality). Indeed, LOCC measurement of quantum states is important for cryptographic protocols Bennett84Ekert91 , channel capacities Hayden04Watrous05 , and distributed quantum information processing Cirac99 . Intuitively we expect that entangled states are more difficult to discriminate locally, since inherently they possess properties that are non-local. Indeed it is known that entanglement can make LOCC discrimination more difficult Terhal01 . But the exact relation is thus far unclear, and there are no general quantitative results. The results that are known can be confusing. One of the earliest results on the subject reveals a set of non-entangled, product states that cannot be discriminated perfectly by LOCC Bennett99 . Later it was shown that any two pure states can be discriminated optimally by LOCC, no matter how entangled they are Walgate00Virmani01 . There have been several results since then on various LOCC settings Chen02-Hillery02 , and connections have been made to bipartite entanglement distillation and formation Badziag03etal . However many results are specific to the bipartite case, or only valid for specific scenarios. We show a clear connection between distancelike measures of entanglement and LOCC state discrimination in the general multipartite case. We first show how the conditions imposed on the measurement by perfect state discrimination can be rewritten in terms of a quantity which looks like a “distance” to the closest separable state. As we weaken these conditions, we then show that this relates directly to three entanglement measures. Finally, combining these results gives a general (pure and mixed state) bound, and, for pure states, allows the following interpretation: entanglement gives an upper bound to the number of pure states that can be discriminated perfectly by LOCC. By using known entanglement results we will give examples of existing and new LOCC discrimination bounds in a unified manner. Theorem 1: A necessary condition for deterministic LOCC discrimination of set $`\{\rho _i|i=1..N\}`$ is that the following inequality holds: $$\underset{i}{}d(\rho _i)D,$$ (1) where $`D`$ is the total dimension of the system, and $`d(\rho _i):=\mathrm{min}{\displaystyle \frac{1}{\mathrm{tr}\{\rho _i\omega _i\}}}`$ such that $`(i)\mathrm{I}1{\displaystyle \frac{\omega _i}{\mathrm{tr}\{\rho _i\omega _i\}}}0,(ii)\omega _iSEP,`$ (2) where $`SEP`$ denotes the set of separable operators. To prove theorem 1, we begin by listing some conditions that the POVMs (positive operator value measures) must satisfy. The task of state discrimination is to perform a measurement (in our case by LOCC) on a system to find out which one of a set of states the system is in. If it is possible to perfectly discriminate among a set of density matrices $`𝒮:=\{\rho _i|i=1..N\}`$ by LOCC, then it is necessary that there exists a POVM $`\{M_i\}`$ satisfying the following conditions: $`{\displaystyle \underset{i}{}}M_i`$ $`=`$ $`\mathrm{I}1`$ (3) $`\mathrm{I}1M_i`$ $``$ $`0`$ (4) $`i\mathrm{tr}\{M_i\rho _i\}`$ $`=`$ $`1`$ (5) $`iM_i`$ $``$ $`SEP`$ (6) Conditions (3) and (4) are simply the conditions mean that $`\{M_i\}`$ is a POVM. Condition (5) says that, given a state $`\rho _i`$, the result corresponding to outcome $`M_i`$ occurs with probability $`1`$, i.e. the discrimination is deterministic. Condition (6) is known to be a necessary condition if the POVM $`\{M_i\}`$ is to be implementable by LOCC Terhal01 . To make the connection to distances between states we first notice that any POVM element $`M_i`$ can be expressed as a positive number $`s_i=\mathrm{tr}\{M_i\}`$ times a density matrix $`\omega _i`$, $`M_i=s_i\omega _i`$. We can then use this to immediately rewrite (3)-(6). Condition (5) is rewritten $`s_i=1/\mathrm{tr}\{\rho _i\omega _i\}`$. Condition (6) means $`\omega _i`$ is separable. For pure states $`s_i`$ now looks like a distancelike quantity between state $`\rho _i`$ and some separable state $`\omega _i`$, such that the remaining conditions are satisfied (that is $`_is_i\omega _i=\mathrm{I}1`$, $`\mathrm{I}1s_i\omega _i0`$). If we then minimise $`s_i`$ such that conditions (4), (5) and (6) are satisfied for each $`i`$ independently, we get exactly the definition of $`d(\rho _i)`$ in theorem 1 (2). Condition (3) implies this minimisation must satisfy $$\underset{i}{}d(\rho _i)D,$$ (7) completing the proof.$`\mathrm{}`$ At this point $`d(\rho )`$ cannot be considered a ‘distance to the closest separable state’ entanglement measure. It turns out that condition (i) in (2) complicates things a lot, and indeed, even without this condition it is not an entanglement monotone for mixed states \[see comment below the definition of the geometric measure (15)\]. Hence the connection to entanglement is not immediate. However, we can use this quantity to relate the problem of state discrimination to other distance-like entanglement monotones, as in the following theorem. Theorem 2: The following bounds hold for all states $`\rho `$: $$d(\rho )r(\rho )2^{E_R(\rho )+S(\rho )}2^{G(\rho )},$$ (8) where $`G(\rho )`$ is the geometric measure; $`E_R(\rho )`$ is the relative entropy of entanglement; $`S(\rho )`$ is the von Neumann entropy; and $`r(\rho ):=|P|(1+R_G(P/|P|)`$, where $`P`$ is the support of state $`\rho `$ NoteSupport , $`|P|:=`$tr$`\{P\}`$, and $`R_G(\rho )`$ is the robustness of entanglement of state $`\rho `$. In the pure state case $`S(\rho )=0`$ and $`P=\rho `$, and so these quantities become exactly (up to log) the geometric measure of entanglement, the relative entropy of entanglement and the robustness of entanglement (from right to left). In the mixed state case, they include some quantification of how mixed the state is. This makes sense in the problem of state discrimination, since the more mixed the states are, the fewer orthogonal states there can be for a given Hilbert space dimension $`D`$. We will later show that the quantities in eq.(8) are equivalent for GHZ sates (these are multipartite states defined originally in GHZ ). To prove the relationship to the robustness of entanglement we must first write $`d(\rho )`$ in a more convenient form. We can rewrite condition $`(i)`$ in (2), as $`\psi |\omega |\psi \mathrm{tr}\{\rho \omega \}|\psi `$. By considering the spectral decomposition of $`\omega `$, it follows that $`\omega `$ can always be rewritten in the form $`\omega =\lambda |P|\frac{P}{|P|}+(1\lambda |P|)\mathrm{\Delta }`$ with the additional conditions $`\mathrm{tr}\{P\mathrm{\Delta }\}=0`$ and $`\lambda \psi |\omega |\psi |\psi `$. We can then rewrite $`d(\rho )=\mathrm{min}(1/\lambda )`$ $`\mathrm{such}\mathrm{that}\mathrm{a}\mathrm{state}\mathrm{\Delta },\mathrm{satisfying}`$ $`\omega =\lambda |P|{\displaystyle \frac{P}{|P|}}+(1\lambda |P|)\mathrm{\Delta }SEP,`$ $`\mathrm{tr}\{P\mathrm{\Delta }\}=0,\lambda \psi |\omega |\psi |\psi .`$ We can now compare this to the global robustness of entanglement $`R_g(\rho )`$ Vidal99 . $`R_g(\rho ):=\mathrm{min}t`$ (10) $`\mathrm{such}\mathrm{that}\mathrm{a}\mathrm{state}\mathrm{\Delta },\mathrm{satisfying}`$ $`{\displaystyle \frac{1}{1+t}}(\rho +t\mathrm{\Delta })SEP.`$ We can understand this as the minimum (arbitrary) noise $`\mathrm{\Delta }`$ that we need to add to make the state separable. We can see that the global robustness of entanglement of the support of state $`\rho `$, $`R_G(P/|P|)`$, is very similar in definition to $`d(\rho )`$ above, (Bounds on Multipartite Entangled Orthogonal State Discrimination Using Local Operations and Classical Communication). The crucial difference being the removal of the two conditions in the last line of (Bounds on Multipartite Entangled Orthogonal State Discrimination Using Local Operations and Classical Communication). Since relaxing conditions can only lead to a lower minimum, we can see that $$d(\rho )r(\rho ):=|P|\left[1+R_g(P/|P|)\right],$$ (11) proving the left inequality of theorem 2. For the centre and right inequalities of theorem 2, we consider the two quantities separately. The relative entropy of entanglement is defined as Vedral98 , $`E_R(\rho ):=\underset{\omega SEP}{\mathrm{min}}S(\rho ||\omega ),`$ (12) where $`S(\rho ||\omega )=S(\rho )\mathrm{tr}\{\rho \mathrm{log}_2\omega \}`$ is the relative entropy and $`S(\rho )`$ is the von Neumann entropy. From the definition of $`R_g(\rho )`$, we know that for some state $`\mathrm{\Delta }`$, the state given by: $`\omega _i:=[P_i/|P_i|+R_g(P_i/|P_i|)\mathrm{\Delta }]/[1+R_g(P_i/|P_i|)]`$ is a separable state. Hence the following inequalities must hold: $`E_R(\rho _i)`$ $`+`$ $`S(\rho _i)`$ (13) $``$ $`\mathrm{tr}\left\{\rho _i\mathrm{log}_2\left({\displaystyle \frac{P_i/|P_i|+R_g(P_i/|P_i|)\mathrm{\Delta }}{1+R_g(P_i/|P_i|)}}\right)\right\}`$ $``$ $`\mathrm{tr}\left\{\rho _i\mathrm{log}_2\left({\displaystyle \frac{P_i/|P_i|}{1+R_g(P_i/|P_i|)}}\right)\right\}`$ $`=`$ $`\mathrm{log}_2\left[|P_i|\left(1+R_g\left(P_i/|P_i|\right)\right)\right],`$ where the third line follows from the monotonicity of the logarithm, which states that $`\mathrm{log}(A+B)\mathrm{log}(A)`$ whenever $`B0`$, for two operators $`A,B`$ Bhatia . The last line is true even if $`\rho _i`$ is any state in the span of $`P_i`$. Hence: $`2^{E_R(\rho _i)+S(\rho _i)}r(\rho _i).`$ (14) We call the geometric measure $`G(\rho )`$ $$G(\rho ):=\mathrm{log}_2\left\{\underset{\omega SEP}{\mathrm{max}}\mathrm{tr}\{\rho \omega \}\right\}.$$ (15) In the case of pure states, this reduces to the geometric measure of entanglement Wei03 . However, for mixed states, this is not an entanglement monotone (for example it is maximised by the maximally mixed state). We immediately see that this would be equivalent (up to log) to $`d(\rho )`$ in (2) if we were to drop condition $`(i)`$. Hence we have $`d(\rho )2^{G(\rho )}`$. However, it is possible to show a stronger bound. In Wei04 it was shown that in the pure state case that $`G(\rho )`$ is bounded from above by the relative entropy of entanglement. We use the same simple concavity arguments now for the mixed state case. By definition $`E_R(\rho _i)+S(\rho _i)=\mathrm{max}_{\omega SEP}\mathrm{tr}\{\rho \mathrm{log}_2\omega \}`$. By concavity of the logarithm, we have for all $`\rho ,\omega `$, $`\mathrm{tr}\{\rho \mathrm{log}_2\omega \}\mathrm{tr}\{\rho \omega \}`$. Thus $`E_R(\rho )+S(\rho )G(\rho )`$ (16) Combining (11), (14) and (16) we get theorem 2. $`\mathrm{}`$ We will now look at how we can use our necessary conditions to bound the maximum number of states that can be discriminated locally. Combining theorem 1 and 2, and dividing by $`N`$, we obtain the following corollary. Corollary: The number of states $`N`$ that can be discriminated perfectly by LOCC is bounded by $`ND/\overline{d(\rho _i)}D/\overline{r(\rho _i)}`$ $``$ $`D/\overline{2^{E_R(\rho _i)+S(\rho _i)}}D/\overline{2^{G(\rho _i)}},`$ where $`\overline{x_i}:=1/N_{i=1}^Nx_i`$, denotes the ‘average’. Hence, in the pure state case, where the bounding quantities reduce to the geometric measure of entanglement, the relative entropy of entanglement and the robustness of entanglement (from right to left), we can interpret these three distance like entanglement measures as bounds on the number of pure states that we can discriminate perfectly by LOCC (see Fig 1). Given this hierarchy of bounds (Bounds on Multipartite Entangled Orthogonal State Discrimination Using Local Operations and Classical Communication), we can apply known results from entanglement theory to find some bounds on $`N`$, one of which we will show is tight. Firstly, the robustness of entanglement is completely solved for pure bipartite states Vidal99 . For a state with Schmidt decomposition $`|\psi =_i\alpha _i|ii`$, the robustness was found to be $`R_g(\psi )=(_i\alpha _i)^21`$. We can immediately put this into (Bounds on Multipartite Entangled Orthogonal State Discrimination Using Local Operations and Classical Communication). For instance, if we have a set of pure bipartite states all with the same entanglement, ($`Bi`$), we have $`N(Bi)d_1d_2/({\displaystyle \underset{i}{}}\alpha _i)^2,`$ (18) where $`d_1,d_2`$ are the dimensions of the Hilbert spaces and $`\alpha _i`$ are the Schmidt coefficients for any one of the states in the set. This has the consequence that it is impossible to distinguish more than $`d`$ maximally entangled states (where $`d`$ is the dimension of one subspace, then $`(\alpha _i)^2=d`$), reproducing a known result Ghosh0104 ; Nathanson04 . In the multiparty case, we know from Wei et al Wei04 that for the $`m`$-party W state $`|W:=|00\mathrm{}01+|\mathrm{00..10}+..+|\mathrm{01..00}+|\mathrm{10..00}`$ and GHZ state $`|GHZ:=|0^m+|1^m`$, the relative entropy of entanglement and the geometric measure coincide and are given by $`E_R(|GHZ)=E_G(|GHZ)=1`$ and $`E_R(|W)=E_G(|W)=\mathrm{log}_2(m/(m1))^{(m1)}`$. Therefore, for any set of states where the state with the average geometric measure (or the lowest) is that of GHZ or W, we have $`N(GHZ)`$ $``$ $`2^{m1}`$ $`N(W)`$ $``$ $`2^m\left((m1)/m\right)^{(m1)}.`$ (19) In fact, if we now call $`N(𝒮_{GHZ})`$ the maximum number of states, in a set made of all GHZ type states (i.e. GHZ up to local unitary transformations), that can be discriminated perfectly by LOCC, then we can show $`N(𝒮_{GHZ})=2^{m1}`$ by explicit construction. We form a set of states $`𝒮_{GHZ}=\{|GHZ_i:=\mathrm{I}1U_i|GHZ\}_{i=1}^{2^{m1}}`$ by local unitaries $`\{U_i\}`$ over $`m1`$ parties. The $`\{U_i\}`$ are formed from all the possible combinations of products of the identity and $`\sigma _x`$ Pauli operations, e.g. $`U_1=\mathrm{I}1^{m2}\sigma _x`$, giving a set of $`_{k=0}^{m1}\left(\genfrac{}{}{0pt}{}{m1}{k}\right)=2^{m1}`$ states. It is easy to check that these can be discriminated by making local $`\sigma _z`$ measurements. Calling $`𝒮_W`$ a set of states equal to the W state up to local unitary transformations, with (Bounds on Multipartite Entangled Orthogonal State Discrimination Using Local Operations and Classical Communication) gives $`N(𝒮_W)<N(𝒮_{GHZ}).`$ (20) We also note that if we can find such a bound by any of the entanglement measures in (Bounds on Multipartite Entangled Orthogonal State Discrimination Using Local Operations and Classical Communication) and show it is tight, those measures below it in the hierarchy are equal. The GHZ case is such an example giving $`R_G(|GHZ)=1`$, and is one of the few cases where the global robustness of entanglement is known for multiparty systems. We round off the examples by showing another simple known result. If even one state in a complete basis is entangled, then (Bounds on Multipartite Entangled Orthogonal State Discrimination Using Local Operations and Classical Communication) shows that the basis cannot be discriminated perfectly Horodecki03 . The simplicity of the basis for the proofs of main results here allows it to be used with other necessary conditions on LOCC measurements. The condition of separability (6), for example, may be changed to more tractable conditions such as positivity of partial transpose or bi-separability Rains99 . It can easily be seen that these conditions would lead to analogous bounds to those derived above. In the case of bi-separability, the example of bipartite states above shows that, for pure states, it is always possible to give some easily computable bound. We have given an interpretation of the global robustness of entanglement, the relative entropy of entanglement and the geometric measure of entanglement as bounds on the number of pure states that can be discriminated perfectly by LOCC. Our general mixed state results imply that the presence of entanglement guarantees a certain minimal level for this difficulty. The difficulty of LOCC state discrimination is an important consideration in various quantum information tasks, (e.g. quantum data hidingTerhal01 ), which may give more uses of these results. This is the topic of ongoing investigations. In this direction, it is also possible to extend theorem 1 to the case of imperfect discrimination. This leads to bounds on the LOCC accessible information, as in Nathanson04 ; DiVincenzo02 , which will be presented in a separate paper. We thank Keiji Matsumoto and Martin Plenio for useful discussions. This work was sponsored by the Asahi Glass Foundation, the JSPS, Leverhulme Trust, and the Royal Commission for the Exhibition of 1851.
warning/0506/cond-mat0506061.html
ar5iv
text
# Critical points and non-Fermi liquids in the underscreened pseudogap Kondo model ## I Introduction Quantum impurity models were introduced in the context of dilute magnetic moments in metals and are based on the notion that important aspects of the physics of extended systems can be captured by spatially local processes. Indeed, the reduction of a complicated many-body problem to the study of a single quantum degree of freedom coupled to a simplified environment can be an important step in understanding more complex physical situations. Impurity models form the basis for the investigation of lattice systems containing local moments – these turn out to be extremely rich, displaying various phenomena such as heavy-electron quantum coherence, glassiness, and quantum phase transitions. In addition, simpler yet interesting quantum effects that take place at the level of a single magnetic moment can also be successfully described using (oversimplified) single impurity models hewson . The fermionic Kondo effect, i.e., the screening of an extra magnetic moment by conduction electrons, is certainly the hallmark of this single-impurity approach, but other aspects like local non-Fermi liquid behavior and boundary quantum phase transitions vojta have also been put forward. In general, single-impurity models can serve as an excellent testing ground for novel techniques and physical paradigms that may be helpful in the more complicated case of extended or dense impurity systems. In this respect, recent work based on the so-called pseudogap (or gapless) Kondo and Anderson models, in which conduction electrons obey a semi-metallic density of states that vanishes as $`\rho (\omega )|\omega |^r`$ ($`r>0`$), has shown interesting promises. Indeed, this problem features a plethora of non-trivial properties, which can be elegantly captured by controlled fermionic renormalization group (RG) techniques fradkin ; fritz1 . Powerful non-perturbative Numerical Renormalization Group (NRG) simulations fritz1 ; ingersent were able to confirm the success of the analytical approach. In particular, recent work fritz1 ; fritz2 has exposed that all fixed points appearing in these pseudogap models can be accessed perturbatively after identifying the value of the DOS exponent $`r`$ that corresponds to their associated lower-critical or upper-critical “dimension”. While the fixed points of the pseudogap spin-1/2 Kondo and single-orbital Anderson models are now understood in this language, here we are interested in certain extensions of these models for which numerical data is also available: the case of a spin-1 impurity has been investigated in the extensive work of Gonzalez-Buxton and Ingersent ingersent , and has shown interesting yet unexplained properties. We note that the physics of spin-1 Kondo models can be realized in multilevel quantum dot systems, and theoretical predictions for transport quantities reflecting the singular Fermi liquid behavior of underscreened Kondo spins in a metallic host singFL have been put forward anna . In this paper, our objective is twofold: first, we will generalize the theories for the critical fixed points, which are already present in the spin-1/2 pseudogap Kondo model, to the $`S=1`$ situation, following Refs. fradkin, ; fritz2, . This includes two types of fixed points, one occuring at particule-hole (p-h) symmetry and being controlled at small $`r`$, and the other present away from p-h symmetry and captured near $`r=1`$ using a general mapping onto an effective interacting resonant level model. Second, we describe the physics of the additional stable fixed point which emerges at particle-hole symmetry, due to the instability of the strong-coupling symmetric fixed point for $`S>1/2`$. Analyzing the leading relevant perturbation, we will be able to capture this non-Fermi phase by a mapping onto an effective ferromagnetic spin $`S_{\mathrm{eff}}=1/2`$ model with singular exponent $`r_{\mathrm{eff}}=r<0`$. Technically, we will perform one-loop perturbative RG calculations together with suitable $`ϵ`$-expansions epsfoot to determine thermodynamic quantities; the results compare favorably with the numerical data obtained in Ref. ingersent, . The present work also illustrates the broad applicability of the general field-theoretic tools, developed in Refs. fradkin, ; fritz1, ; fritz2, , in a more complicated setting. The bulk of the paper is organized as follows: we start by presenting the spin-1 Kondo model with a given pseudogap density of states (Sec. II), and summarize previous results from NRG calculations (Sec. III). In Sec. IV, we focus on the case of particle-hole symmetry. It shows a critical point accessible from weak coupling (which is a generalization to $`S>1/2`$ of the critical point analyzed by Withoff and Fradkin fradkin ). We also present an analysis and strong-coupling calculations for the stable non-Fermi liquid fixed point found in the numerical simulations of Ref. ingersent, . Sec. V presents a novel field theory that describes the quantum phase transition between the free-moment and strong-coupling phases in the presence of particle-hole asymmetry. In all cases, direct comparison to NRG data is made. Technical details on the diagrammatic approach can be found in several appendices. We will finally conclude the paper and discuss some open questions related to the pseudogap Kondo model. ## II The model The pseudogap Kondo model (for a review, see Ref. vojta, ) originates from the question of how a magnetic moment reacts to the coupling to an electronic bath which has depleted weight at the Fermi level, and is of interest to various condensed-matter systems such as $`d`$-wave superconductors, graphite, and semiconductors. The basic physical idea is that the possibility of observing a Kondo effect is weakened by the lack of low-energy states, which results in various quantum phase transitions controlled by the strength of the Kondo interaction, the amount of particle-hole symmetry breaking, and obviously the shape of the electronic density of states. We will model the latter by the simplified form $$\rho (ϵ)=\underset{k}{}\delta (ϵϵ_k)=N_0|ϵ|^r\theta (D^2ϵ^2),$$ (1) which has an algebraic dependence in energy characterized by the number $`r>0`$ (the standard Kondo model corresponds to $`r=0`$). Here $`D`$ is the half-bandwidth and $`N_0=(r+1)/(2D^{r+1})`$. The Kondo model is characterized by an antiferromagnetic coupling $`J_B`$ to this fermionic bath, and possibly a potential scattering term $`E_0`$. It reads $`_B`$ $`=`$ $`{\displaystyle \underset{k\sigma }{}}ϵ_kc_{k\sigma }^{}c_{k\sigma }^{}+J_B\stackrel{}{S}_B{\displaystyle \underset{\sigma \sigma ^{}}{}}c_\sigma ^{}(\mathrm{𝟎}){\displaystyle \frac{\stackrel{}{\tau }_{\sigma \sigma ^{}}}{2}}c_\sigma ^{}^{}(\mathrm{𝟎})`$ (2) $`+E_0{\displaystyle \underset{\sigma }{}}c_\sigma ^{}(\mathrm{𝟎})c_\sigma ^{}(\mathrm{𝟎})`$ in standard notation, and the subscript <sub>B</sub> stands for “bare” physical quantities and operators, in view of the later RG calculations. For $`E_0=0`$ the model is p-h symmetric, and $`E_0`$ tunes the amount of p-h asymmetry. In the following we will be concerned with the case of a spin-1 operator $`\stackrel{}{S}_B`$ and want to analyze the various fixed points of this model. ## III Summary of previous numerical results The comprehensive numerical study of Gonzalez-Buxton and Ingersent ingersent has exposed that the model (1-2) displays several zero-temperature phases, separated by boundary quantum phase transitions. Their results are summarized in Fig. 1 as a function of Kondo coupling $`J_B`$, particle-hole asymmetry $`E_0`$, and the value of the bath exponent $`r`$. Let us focus on the p-h symmetric case, $`E_0=0`$, first. At $`J_B=0`$ we have a free spin $`1`$, dubbed local-moment (LM) fixed point, which is stable w.r.t a small Kondo coupling $`J_B`$ for $`r>0`$. Flow towards large coupling is possible only when $`J_B`$ exceeds a critical value $`J_c`$ – this phase transition defines a symmetric critical (SCR) point. Contrary to the $`S=1/2`$ case where a symmetric strong-coupling (SSC) fixed point (corresponding to $`J_B=\mathrm{}`$) is always attained when $`J_B>J_c`$, SSC is generically unstable in the $`S=1`$ model, which instead shows an intermediate stable fixed point (NFL). It is located at a finite value $`J^{}>J_c`$ and corresponds to a non-Fermi liquid (NFL) phase. This flow structure does not persist for all values of $`r`$, since at a value $`r_{\mathrm{max}}0.27`$, the SCR and NFL fixed points collapse and disappear altogether. For $`r>r_{\mathrm{max}}`$, LM becomes then generically stable rmaxfoot . Now we turn to finite p-h asymmetry, i.e., potential scattering $`E_00`$. Then SCR is stable against finite $`E_0`$ only for $`0<r<r^{}`$ (where $`r^{}0.245`$), whereas NFL is always unstable. In the case $`r>r^{}`$, SCR becomes unstable as well, so that an asymmetric cousin of the critical point, ACR, exists and controls the quantum phase transition between LM and an asymmetric strong-coupling (ASC) fixed point. ACR is associated to finite values $`J_c^{}`$ and $`E_{0c}`$ of both the Kondo coupling and the potential scattering term. The numerical simulations of Ref. ingersent, have also established the behavior of several observables, such as the impurity contributions to the entropy and to the Curie susceptibility, for the various fixed points. Whereas the trivial cases (LM, SSC, ASC) are readily understood, all intermediate-coupling fixed points (SCR, NFL, ACR) correspond to fractional values of both the ground state degeneracy and spin, and cannot be understood on the basis of an independent electron picture. The values extracted from the Numerical Renormalization Group data are shown on Table I, and in the remainder of the paper we aim at an analytical calculation of such singular properties, which characterize the complicated nature of the various ground states occuring in the model. ## IV Particle-hole symmetric model In this section we analyze the particle-hole symmetric spin-1 pseudogap Kondo model (2), with $`E_0=0`$. We start with the weak-coupling RG which allows to capture the SCR fixed point for small $`r`$. Then we consider an expansion around the symmetric strong-coupling fixed point – this will lead us to an effective ferromagnetic Kondo model with singular density of states which we use to describe the physics of the stable NFL fixed point. ### IV.1 Weak-coupling analysis of the SCR fixed point The study of the regime with small Kondo coupling and particle-hole symmetry was done by Withoff and Fradkin fradkin for $`S=1/2`$. Perturbative RG is performed around $`J_B=0`$, i.e., the LM fixed point. The tree level scaling dimension of the Kondo coupling is $`\mathrm{dim}[J_B]=r`$, and the RG flow in the vicinity of the LM fixed point can be described by the one-loop beta function: $$\beta (j)=rjj^2+𝒪(j^3)$$ (3) where $`j`$ is the renormalized dimensionless Kondo coupling, defined in Eq. 7 below. The beta function (3) obviously yields an unstable fixed point at $`j_c=r`$, which corresponds to SCR and controls the transition between LM and NFL. The correlation length exponent is $$\frac{1}{\nu }=r+𝒪(r^2),$$ (4) and $`r=0`$ can be interpreted as lower-critical “dimension” for the phase transition. These results can be expected to persist for $`S=1`$, since the beta function for the Kondo problem is known fabrizio to be independent of the impurity spin value $`S`$. However, as we are interested in computing physical observables at SCR, which do differ from their $`S=1/2`$ counterparts, we will proceed with a complete RG analysis for $`S=1`$ as well. We only sketch here the intermediate results, while details on the derivation can be found in Appendix A. #### IV.1.1 RG procedure The analysis is based on a representation of the spin 1 by a triplet of Abrikosov fermions, $`f_{mB}`$: $`S_B^+`$ $`=`$ $`S_B^x+iS_B^y=\sqrt{2}(f_{1B}^{}f_{0B}^{}+f_{0B}^{}f_{1B}^{}),`$ $`S_B^z`$ $`=`$ $`f_{1B}^{}f_{1B}^{}f_{1B}^{}f_{1B}^{}`$ (5) where the Hilbert space constraint $$Q=\underset{m=1,0,1}{}f_{mB}^{}f_{mB}^{}=1$$ (6) is enforced kircan ; lambda using a chemical potential $`\lambda \mathrm{}`$. To proceed with the RG scheme, we relate the bare coupling constant and field, $`J_B`$ and $`f_{mB}`$, to renormalized quantities, $`j`$ and $`f_m`$ (which bear no index to shorten the notation within renormalized perturbation theory): $`N_0J_B`$ $`=`$ $`\mu ^rZ_f^1Z_Jj,`$ (7) $`f_{mB}`$ $`=`$ $`Z_f^{1/2}f_m^{}`$ (8) where $`\mu `$ is a renormalization energy scale. These equations define renormalization factors $`Z_J`$ and $`Z_f`$ which will be determined using the standard field-theoretic RG scheme bgz , employing dimensional regularization and minimal subtraction of poles. A one-loop calculation (see Appendix A) of the $`f_m^{}`$ self-energy and the Kondo vertex yields $`Z_f`$ $`=`$ $`1+𝒪(j^2),`$ (9) $`Z_J`$ $`=`$ $`1+{\displaystyle \frac{j}{r}}+𝒪(j^2),`$ (10) respectively. The RG beta function is obtained by the condition that the bare Kondo coupling is scale invariant, i.e. $`dJ_B/d\mu =0`$, which gives from (7): $$\beta (j)\mu \frac{dj}{d\mu }=\frac{rj}{1+j\frac{d}{dj}\mathrm{ln}(Z_J/Z_f)}.$$ (11) Inserting (9)-(10) into this expression proves that the beta function in the $`S=1`$ case is still given by Eq. (3), as expected. We also note that p-h asymmetry is irrelevant at this fixed point, with $`\beta (e_0)=re_0`$ (where $`e_0N_0E_0`$ is the dimensionless potential scattering term). #### IV.1.2 Impurity entropy The impurity contribution to the entropy is obtained as the difference of the total entropy and the entropy of the electron bath alone. Its $`T=0`$ value represents a measure of the impurity ground state degeneracy. For the SCR fixed point it turns out that corrections to the LM value are extremely small kircan : $$S_{\mathrm{imp}}^{\mathrm{SCR}}=\mathrm{ln}3+𝒪(r^3).$$ (12) This makes the comparison to numerical data difficult, and we refrain from extracting the prefactor of the $`r^3`$ term. #### IV.1.3 Impurity susceptibility The impurity susceptibility $`\chi _{\mathrm{imp}}`$ is similarly calculated by removing the free (i.e. $`J_B=0`$) contribution of the bulk fermions to the total spin susceptibility suscfoot , $$\chi _{\mathrm{tot}}(T)=_0^\beta d\tau <S_{\mathrm{tot}}^z(\tau )S_{\mathrm{tot}}^z(0)>,$$ (13) with $$S_{\mathrm{tot}}^z=\underset{m}{}mf_{mB}^{}f_{mB}^{}+\underset{𝐱\sigma }{}\frac{(1)^\sigma }{2}c_\sigma ^{}(𝐱)c_\sigma ^{}(𝐱).$$ (14) Apart from numerical factors, the calculation is analogous to the $`S=1/2`$ case kircan , and we find at lowest order in $`j`$: $$\chi _{\mathrm{imp}}=\frac{2}{3T}j\frac{2}{3T}\left(\frac{\mu }{T}\right)^r_{\mathrm{}}^+\mathrm{}dx\frac{e^x}{(e^x+1)^2}$$ (15) Because we want to obtain the SCR fixed point contribution, we can replace $`j=j_c=r`$ and take $`(\mu /T)^r1`$ as $`r0`$, which gives finally the effective Curie constant: $$T\chi _{\mathrm{imp}}^{\mathrm{SCR}}=\frac{2}{3}\frac{2r}{3},$$ (16) in agreement with the numerical value found in Table I. ### IV.2 Non-Fermi liquid phase near strong coupling #### IV.2.1 Around the limit $`J=\mathrm{}`$ To understand the emergence of a stable non-trivial fixed point, we consider in the first place the vicinity of the SSC fixed point. Indeed, the renormalization flow sketched from the numerical solution demonstrates that the strong-coupling fixed point is unstable, which is quite curious at first thought. Indeed, at large $`J`$, the spin 1 is partially screened by the fermions at site $`\mathrm{𝟎}`$, leaving a remnant effective spin $`S_{\mathrm{eff}}=1/2`$. Applying the Nozières-Blandin argument nozieres we see that an effective ferromagnetic coupling $`J_{\mathrm{eff}}=t_{01}^2/J`$ is then generated between $`\stackrel{}{S}_{\mathrm{eff}}`$ and the electrons $`c_\sigma ^{}(\mathrm{𝟏})`$ sitting on the site near $`c_\sigma ^{}(\mathrm{𝟎})`$ (above, $`t_{01}`$ is the typical hopping between $`c_\sigma ^{}(\mathrm{𝟎})`$ and $`c_\sigma ^{}(\mathrm{𝟏})`$, and we assumed a chain representation of the electronic bath). Naively we would then expect SSC to be stable. This is not the case, because at particle-hole symmetry the Green’s functions of the above electrons are related through: $$<c_\sigma ^{}(\mathrm{𝟎},i\omega )c_\sigma ^{}(\mathrm{𝟎},i\omega )>^1=i\omega t_{01}^2<c_\sigma ^{}(\mathrm{𝟏},i\omega )c_\sigma ^{}(\mathrm{𝟏},i\omega )>$$ (17) which shows that the fermions $`c_\sigma ^{}(\mathrm{𝟏})`$ have an effective density of states $`\rho _{\mathrm{eff}}(ϵ)|ϵ|^{r_{\mathrm{eff}}}`$, with $`r_{\mathrm{eff}}=r<0`$. Thus, the effective Hamiltonian near SSC is a ferromagnetic spin-1/2 Kondo model with a negative DOS exponent. Interestingly, this model was studied before in Ref. bulla, , and was found to display a stable intermediate-coupling fixed point. To see this, we note that the beta function for $`J_{\mathrm{eff}}`$ reads $$\beta (j_{\mathrm{eff}})=r_{\mathrm{eff}}j_{\mathrm{eff}}j_{\mathrm{eff}}^2=rj_{\mathrm{eff}}j_{\mathrm{eff}}^2.$$ (18) This yields a stable fixed point at $`j_{\mathrm{eff}}^{}=r`$, which is allowed since the effective coupling is ferromagnetic! In our original language, this is the NFL fixed point, which controls now a whole phase with non-trivial properties, and is a unique feature of the $`S>1/2`$ pseudogap Kondo model. In particular, the above arguments show that this fixed point is located at $`j^{}1/r`$. The NFL fixed point is unstable with respect to particle-hole asymmetry, with $`\beta (e_0)=re_0`$, a situation which will be discussed in Sec. V. Finally, we note that the values for $`r^{}`$ and $`r_{\mathrm{max}}`$ obtained by the Numerical Renormalization Group for the original problem ingersent and for the effective model bulla coincide, in agreement with our identification. We would like now to compute the physical quantities at the intermediate NFL fixed point. #### IV.2.2 Impurity entropy We know that the correction to the entropy at a particle-hole symmetric Withoff-Fradkin type of fixed point is as small as $`r^3`$. However, we are now expanding around a strong-coupling fixed point and not with respect to a free spin $`S=1`$. Since the impurity entropy is defined relative to the decoupled ($`J=0`$) limit, we have to take into account the contribution at SSC ($`J_B=\mathrm{}`$), which contains a $`\mathrm{ln}2`$ term due to the underscreened spin $`S_{\mathrm{eff}}=1/2`$, and depends on $`r`$ as well ingersent ; fritz1 , due to the singular nature of the resonant level limit (although the SSC fixed point is trivial). We obtain finally $$S_{\mathrm{imp}}^{\mathrm{NFL}}=\mathrm{ln}2+2r\mathrm{ln}2+𝒪(r^3),$$ (19) as anticipated numerically (Table I). #### IV.2.3 Impurity susceptibility We have to follow the same argument as above, and sum the contribution of the strong-coupling fixed point, $`T\chi _{\mathrm{imp}}^{\mathrm{SSC}}=1/4+r/8`$ (see fritz1 ) as well as the non-trivial correction due to the stable $`S_{\mathrm{eff}}=1/2`$ ferromagnetic fixed point with exponent $`r_{\mathrm{eff}}=r`$, which is kircan $`\mathrm{\Delta }(T\chi _{\mathrm{imp}})=r_{\mathrm{eff}}/4=r/4`$. The complete result thus agrees with the value in Table I: $$T\chi _{\mathrm{imp}}^{\mathrm{NFL}}=\frac{1}{4}+\frac{3r}{8}.$$ (20) ## V Particle-hole asymmetric model The nature of the particle-hole asymmetric fixed point of the pseudogap Kondo model, present at $`E_00`$ and $`r>r^{}`$, was understood in the recent work fritz1 ; fritz2 . The crucial point noticed there is that this quantum phase transition separates the local-moment fixed point from the asymmetric strong-coupling one, as can be checked in Fig. 1, so that the corresponding field theory can be constructed by “mixing” the relevant degrees of freedom associated to both fixed points. In the $`S=1/2`$ case, one needed to consider a level crossing of a doublet (associated to LM) and a singlet (from to ASC), with transitions allowed through the coupling to the conduction electrons. Not surprisingly, the effective theory took the form of an infinite-$`U`$, i.e., maximally p-h asymmetric, Anderson model. This was perfectly consistent with numerical data which indicate the phase transitions in the pseudogap Kondo and Anderson models are in the same universality class. The theory could be analyzed by expanding around the level-crossing (valence fluctuation) fixed point, and it was shown that $`r=1`$ plays the role of an upper-critical “dimension”. In the following, we want to generalize the idea of Refs. fritz1, ; fritz2, to the spin-1 case. ### V.1 Derivation of the critical field theory In the present problem, the spin $`S=1`$ will be described by a triplet of fermions $`f_m^{}`$ near LM. On the other hand, ASC shows underscreening of the spin, and gives a remnant spin-1/2 moment, that we can describe by a doublet of bosons $`b_\sigma ^{}`$. Having now the important variables to describe the transition at ACR, we need to write down the effective theory. Several requirements have to be imposed: the Hamiltonian should obey spin rotation invariance and should reduce to a spin 1 (resp. 1/2) Kondo Hamiltonian near the LM (resp. ASC) limit. One naturally arrives at the following generalization of the infinite-$`U`$ Anderson model of Ref. fritz1, : $``$ $`=`$ $`{\displaystyle \underset{k\sigma }{}}ϵ_kc_{k\sigma }^{}c_{k\sigma }^{}+ϵ_f{\displaystyle \underset{m}{}}f_m^{}f_m^{}`$ $`+`$ $`V_B\left[f_{0B}^{}b_B^{}c_{}^{}(\mathrm{𝟎})+f_{0B}^{}b_B^{}c_{}^{}(\mathrm{𝟎})\right]+\mathrm{h}.\mathrm{c}.`$ $`+`$ $`\sqrt{2}V_B\left[f_{1B}^{}b_B^{}c_{}^{}(\mathrm{𝟎})+f_{1B}^{}b_B^{}c_{}^{}(\mathrm{𝟎})\right]+\mathrm{h}.\mathrm{c}..`$ Here, $`ϵ_f`$ is the parameter used to tune the system through the phase transition; at tree level the transition occurs at $`ϵ_f=0`$, and corresponds to the crossing of doublet and triplet states, which we call the valence fluctuation (VFl) fixed point. The restriction to the physical Hilbert space is implemented with the constraint $$Q=\underset{m}{}f_{mB}^{}f_{mB}^{}+\underset{\sigma }{}b_{\sigma B}^{}b_{\sigma B}^{}=1.$$ (22) One can check easily via a Schrieffer-Wolff transformation that $``$ (V.1) evolves onto a spin 1 (resp. 1/2) Kondo model when $`ϵ_f`$ is large and negative (resp. positive), as it should be. The effective Hamiltonian (V.1) seems to have nothing to do with the original model (2), and in particular it is not as simply understandable as in the $`S=1/2`$ case considered in fritz2 . However, we will show that it correctly describes the critical properties at ACR. A number of facts are encouraging: Power counting at the $`ϵ_f=V_B=0`$ fixed point shows that the coupling constant $`V_B`$ has a tree-level scaling dimension $`\mathrm{dim}[V_B]=\overline{r}=(1r)/2`$, i.e., it is relevant for $`r<1`$, which indicates the possibility of a non-trivial critical point in this regime. In contrast, $`V_B`$ is irrelevant and will flow to zero for $`r>1`$, which establishes the role of $`r=1`$ as upper-critical dimension where $`V_B`$ is marginal. Similar to the situation in the spin-1/2 model, the transition will become a level crossing with perturbative corrections for $`r>1`$. For $`V_B0`$ we have strong “valence” fluctuations between the doublet and triplet levels, leading to an entropy of $`\mathrm{ln}5`$, in agreement with the numerical result shown in Table I. Motivated by this we proceed with an RG expansion around the decoupled $`V_B=0`$ fixed point. As demonstrated below, we find a flow diagram identical in structure to the one of the spin-1/2 model, Fig. 1 of Ref. fritz2, , with the difference that the VFl fixed point now corresponds to a crossing of doublet and triplet levels. This flow diagram is very similar to the one of a standard $`\varphi ^4`$ theory. In particular, the non-trivial critical point occuring for $`r<1`$ (which can be understood as the analogue of the Wilson-Fisher fixed point) will be identified with the asymmetric critical point (ACR) of the original Kondo Hamiltonian. ### V.2 RG calculation for $`r`$ close to 1 #### V.2.1 RG procedure The renormalization method developed in Ref. fritz2, differs slightly from the technique used in Sec. IV.1 and Appendix A. We still need to introduce renormalization factors for the hybridization strength and fields: $`N_0V_B`$ $`=`$ $`\mu ^{+\overline{r}}Z_f^{1/2}Z_b^{1/2}Z_Vv,`$ (23) $`f_{mB}`$ $`=`$ $`Z_f^{1/2}f_m^{},`$ (24) $`b_{\sigma B}`$ $`=`$ $`Z_b^{1/2}b_\sigma ^{},`$ (25) as well as mass renormalization terms, $$_\lambda =\delta \lambda _f\underset{m}{}f_m^{}f_m^{}+\delta \lambda _b\underset{\sigma }{}b_\sigma ^{}b_\sigma ^{},$$ (26) already written using renormalized quantities. We have introduced again a running renormalization scale $`\mu `$ and used $`\overline{r}=(1r)/2`$. The RG will be performed at criticality, i.e., the $`\delta \lambda `$ are chosen such to cancel the real parts of the bare self-energies. Including the contribution of the counter-terms associated to the above renormalization constants, we find (see Appendix B) the $`f`$-electron self-energy contribution at one-loop: $$\mathrm{\Sigma }_f(i\omega )=2v^2\left(\frac{\mu }{D}\right)^{\overline{r}}\left[D+\frac{i\omega }{2\overline{r}}\left(\frac{D}{|\omega |}\right)^{\overline{r}}\right]+\delta \lambda _fi\omega (Z_f1).$$ (27) By definition, the field renormalization is determined by the condition $`(/\omega )\mathrm{\Sigma }_f(0)=0`$, and criticality is ensured by the vanishing of the mass, $`\mathrm{\Sigma }_f(0)=0`$, which gives: $`Z_f`$ $`=`$ $`1{\displaystyle \frac{v^2}{\overline{r}}}+𝒪(v^4),`$ (28) $`\delta \lambda _f`$ $`=`$ $`2v^2\left({\displaystyle \frac{\mu }{D}}\right)^{\overline{r}}D+𝒪(v^2).`$ (29) The bosonic self-energy is similarly calculated (only numerical prefactors differ) and leads to: $`Z_b`$ $`=`$ $`1{\displaystyle \frac{3v^2}{2\overline{r}}}+𝒪(v^4),`$ (30) $`\delta \lambda _b`$ $`=`$ $`3v^2\left({\displaystyle \frac{\mu }{D}}\right)^{\overline{r}}D+𝒪(v^2).`$ (31) At this order, the hybridization is left unrenormalized, $`Z_V=1`$. Finally, imposing the scale invariance condition $`\mathrm{d}V_B/\mathrm{d}\mu =0`$, we find the beta function: $`\beta (v)`$ $``$ $`\mu {\displaystyle \frac{\mathrm{d}v}{\mathrm{d}\mu }}={\displaystyle \frac{\overline{r}v}{1+v\frac{d}{dv}\mathrm{ln}(Z_VZ_f^{1/2}Z_b^{1/2})}}`$ (32) $`=`$ $`\overline{r}v+{\displaystyle \frac{5}{2}}v^3+𝒪(v^5).`$ This gives the fixed point value $`(v^{})^2=(2/5)\overline{r}`$. The correlation length exponent of the LM–ASC transition can be obtained similar to Ref. fritz2, , with the result $$\frac{1}{\nu }=r+𝒪(\overline{r}^2)(r<1),$$ (33) whereas a level crossing (formally $`\nu =1`$) occurs for $`r>1`$. The RG flow is thus identical to Fig. 1 of Ref. fritz2, . #### V.2.2 Impurity entropy The correction at order $`v^2`$ to the free energy is evaluated in Appendix B and reads: $$\mathrm{\Delta }F=\frac{12}{5}v^2\mu ^{2\overline{r}}_0^Ddϵ\frac{ϵ^r}{ϵ}\mathrm{tanh}\frac{ϵ}{2T}$$ (34) Using $`\mathrm{\Delta }S_{\mathrm{imp}}=(\mathrm{\Delta }F)/T`$, and taking the limit $`\overline{r}0`$, we obtain finally: $$S_{\mathrm{imp}}^{\mathrm{ACR}}=\mathrm{ln}5\frac{24\mathrm{ln}2}{25}(1r)$$ (35) which is a agreement with the result in Table I. #### V.2.3 Impurity susceptibility The computation of the Curie constant is done along the same line, but is more involved due to the proliferation of Feynman graphs, since the total spin operator $`S_{\mathrm{tot}}^z`$ $`=`$ $`{\displaystyle \underset{m}{}}mf_{mB}^{}f_{mB}^{}`$ (36) $`+`$ $`{\displaystyle \underset{\sigma }{}}{\displaystyle \frac{(1)^\sigma }{2}}\left[b_{\sigma B}^{}b_{\sigma B}^{}+{\displaystyle \underset{𝐱}{}}c_\sigma ^{}(𝐱)c_\sigma ^{}(𝐱)\right]`$ mixes all three possible fields in the calculation. We simply quote the final result, derived in Appendix B: $`\chi _{\mathrm{imp}}`$ $`=`$ $`{\displaystyle \frac{1}{2T}}{\displaystyle \frac{v^2\mu ^{2\overline{r}}}{10T}}{\displaystyle _0^D}\mathrm{d}ϵ{\displaystyle \frac{ϵ^r}{T^2}}[{\displaystyle \frac{\mathrm{sinh}ϵ/T}{1+\mathrm{cosh}ϵ/T}}1`$ (37) $`+{\displaystyle \frac{6\mathrm{sinh}ϵ/T}{(1+\mathrm{cosh}ϵ/T)^2}}].`$ We finally insert the fixed-point value $`v^{}`$ and take the limit $`\overline{r}0`$, which gives: $$T\chi _{\mathrm{imp}}^{\mathrm{SCR}}=\frac{1}{2}\frac{32\mathrm{ln}2}{50}(1r).$$ (38) Note that the leading Curie constant above is associated to the trivial susceptibility of a magnetic system that consists of a triplet and a singlet. Because of numerical prefactors, e.g. in the partition function, it is not simply given by the expression valid for a triplet state only, namely S(S+1)/3=2/3, but by the different value 1/2. Finally, we check that equation (38) coincides with the results in Table I and vindicates the usefulness of the effective theory (V.1) in describing the ACR fixed point. ## VI Conclusion In this paper, we have studied the spin-1 pseudogap Kondo model using perturbative RG together with renormalized perturbation theory. We have shown that all non-trivial fixed points are accessible in the vicinity of their associated upper-critical or lower-critical dimension. In particular, we have proposed a novel field theory governing the critical behavior in the presence of particle-hole asymmetry, and we also have provided a physical picture for the stable particle-hole symmetric non-Fermi liquid phase (which is not present in the $`S=1/2`$ case) using a strong-coupling mapping to an effective ferromagnetic spin-1/2 model with singular density of states. The generic instability of the symmetric strong-coupling fixed point SSC and the emergence of the non-Fermi liquid phase presents a nice example of a situation where the purely local strong-coupling picture of Kondo singlet formation is clearly inappropriate. We also note that, from the arguments given above, our qualitative results are expected to be generic to the pseudogap Kondo model for all spin values $`S`$ greater or equal to 1. The collected analytical results for thermodynamic quantities established in this paper can finally be compared to the numerical values obtained from the simulations ingersent , and are displayed in Fig. 2 and 3. The agreement is excellent, even at one-loop order, demonstrating the usefulness of $`ϵ`$-expansions for impurity models with power-law bath spectra. Using the present RG methods other observables can be calculated as well fritz2 , like the critical exponents for the local susceptibility or the conduction electron $`T`$ matrix. We only mention here the result for the $`T`$ matrix: similar to the spin-1/2 model the critical exponent can be determined exactly to all orders in perturbation theory, with the result $$\mathrm{Im}T(\omega )\frac{1}{\omega ^r}(0<r<1)$$ (39) which holds at all intermediate coupling fixed points (SCR, NFL, ACR). At the critical dimensions, $`r=0`$ and $`r=1`$, logarithmic corrections to this result occur. The application of our results to experiments requires host systems with pseudogap DOS. Candidates are $`d`$-wave superconductors or materials with a special semiconducting band structure (as planar graphite) – these obey a linearly vanishing density of states. Systems with exponent $`0<r<1`$ are difficult to realize; proposals have been made e.g. using disordered mesoscopic conductors hopk . On the theoretical side, the single-impurity pseudogap Kondo problems are now reasonably well understood. One open issue is a possible analytical description of a collapse of intermediate-coupling fixed points, which occurs both as $`rr_{}^{}{}_{}{}^{+}`$ (where two critical fixed points merge) and at $`rr_{\mathrm{max}}^{}{}_{}{}^{}`$ (where the stable and the critical p-h symmetric fixed points meet and disappear). One idea comes to mind when comparing the entropies for a spin-$`S`$ model: the critical fixed point has $`S_{\mathrm{imp}}(J_c)=\mathrm{ln}(2S+1)+𝒪(r^3)`$, whereas for the stable NFL fixed point has $`S_{\mathrm{imp}}(J^{})=\mathrm{ln}(2S)+2r\mathrm{ln}2+𝒪(r^3)`$. The two values cross at $`r_{\mathrm{max}}1/(S\mathrm{ln}8)`$, a value that becomes small if $`S1`$. One may hope that a controlled weak-coupling theory describing the physics of $`r_{\mathrm{max}}`$ or $`r^{}`$ exists for large $`S`$, although this remains to be demonstrated. Finally, we point out that generalizations of the pseudogap model to several impurities can show an interesting interplay of critical points with different nature. ###### Acknowledgements. We thank L. Fritz and M. Kirćan for useful discussions. This research was supported by the DFG Center for Functional Nanostructures and the Virtual Quantum Phase Transitions Institute in Karlsruhe. ## Appendix A Weak-coupling RG Here we provide calculational details for the weak-coupling RG of Sec. IV.1. ### A.1 Vertex renormalization at one loop We first rewrite the bare Kondo Hamiltonian (2) in terms of renormalized quantities, allowing to separate the bare action $`𝒮_B`$ into a renormalized Kondo part $`𝒮`$, a constraint term $`𝒮_\lambda `$, and a series of counter-terms $`\delta 𝒮`$: $`𝒮_B`$ $`=`$ $`𝒮+𝒮_\lambda +\delta 𝒮,`$ (40) $`𝒮`$ $`=`$ $`{\displaystyle _0^\beta }d\tau {\displaystyle \underset{k}{}}\overline{c}_{k\sigma }^{}(_\tau +ϵ_k)c_{k\sigma }^{}+{\displaystyle \underset{m}{}}\overline{f}_m^{}_\tau f_m^{}`$ (41) $`+{\displaystyle _0^\beta }d\tau {\displaystyle \underset{mm^{}\sigma \sigma ^{}}{}}\mathrm{\Gamma }_{mm^{}\sigma \sigma ^{}}^{(0)}\overline{f}_m^{}f_m^{}\overline{c}_\sigma ^{}(\mathrm{𝟎})c_\sigma ^{}(\mathrm{𝟎}),`$ $`𝒮_\lambda `$ $`=`$ $`{\displaystyle _0^\beta }d\tau \lambda \left({\displaystyle \underset{m}{}}\overline{f}_m^{}f_m^{}1\right),`$ (42) $`\delta 𝒮`$ $`=`$ $`{\displaystyle _0^\beta }d\tau \delta \lambda {\displaystyle \underset{m}{}}\overline{f}_m^{}f_m^{}`$ $`+{\displaystyle _0^\beta }d\tau (Z_J1){\displaystyle \underset{mm^{}\sigma \sigma ^{}}{}}\mathrm{\Gamma }_{mm^{}\sigma \sigma ^{}}^{(0)}\overline{f}_m^{}f_m^{}\overline{c}_\sigma ^{}(\mathrm{𝟎})c_\sigma ^{}(\mathrm{𝟎}).`$ With a renormalized dimensionful coupling $`J=j/N_0`$ the tree level vertices satisfy $`\mathrm{\Gamma }_{10}^{(0)}=\mathrm{\Gamma }_{01}^{(0)}=\mathrm{\Gamma }_{01}^{(0)}=\mathrm{\Gamma }_{10}^{(0)}=\mu ^rJ/\sqrt{2}`$ and $`\mathrm{\Gamma }_{mm\sigma \sigma }^{(0)}=\mu ^rJm(1)^\sigma /2`$, with all other terms being zero. Indeed, inserting the definitions (7)-(8) into the above expression for the action, one recovers the bare action of the problem. The free propagators are given by: $`G_{c0}(i\omega ,k)`$ $`=`$ $`{\displaystyle \frac{1}{i\omega ϵ_k}}`$ (44) $`G_{f0}(i\omega )`$ $`=`$ $`{\displaystyle \frac{1}{i\omega \lambda }}`$ (45) where the limit $`\lambda \mathrm{}`$ should be taken only at the end of the calculation lambda . The diagrammatic expansion at one loop of the vertex $`\mathrm{\Gamma }_{10}`$ is given in Fig. 4. Evaluating the loop at zero temperature, one finds: $`\mathrm{\Gamma }_{10}^{(1)}(i\nu )`$ $`=`$ $`\mu ^rZ_J{\displaystyle \frac{J}{\sqrt{2}}}\mu ^{2r}{\displaystyle \frac{J^2}{\sqrt{2}}}{\displaystyle _D^0}dϵN_0{\displaystyle \frac{|ϵ|^r}{ϵ+i\nu }}`$ (46) $`=`$ $`{\displaystyle \frac{J}{\sqrt{2}}}\mu ^r\left[Z_J{\displaystyle \frac{j}{r}}(1+r\mathrm{ln}(|\nu |/\mu ))\right]`$ where we have used $`j=N_0J`$, and the limit $`r0`$ has been taken in the last integral. The renormalization condition for the vertex is $`\mathrm{\Gamma }_{10}^{(1)}(0)=\mathrm{\Gamma }_{10}^{(0)}=\mu ^rJ/\sqrt{2}`$, so that the result (10) follows. ### A.2 Impurity susceptibility The diagrammatic expansion of the impurity susceptibility follows from its definition (13-14), and accounting for all numerical prefactors (which arise either from the $`(1)^\sigma /2`$ term in (14) or from combinatorial reasons), we have to calculate the contribution shown in Fig. 5. Note that $`<Q>\chi _{\mathrm{imp}}`$ and not $`\chi _{\mathrm{imp}}`$ has a simple diagrammatic expansion, since the constraint $`Q=_mf_{mB}^{}f_{mB}^{}=1`$ means that we are not expanding around a free-particle limit. At lowest order, we have $`<Q>=3\mathrm{exp}(\beta \lambda )`$. Computing the Matsubara sums at finite temperature, one gets the result (15) in the limit where the half-bandwidth $`D`$ goes to infinity. ## Appendix B RG around the valence-fluctuation limit This appendix contains details for the RG calculation of Sec. V.2. ### B.1 Pseudoparticle self-energies at one loop Again, we will set up notations and express the bare action $`𝒮_B=𝒮+𝒮_\lambda +\delta 𝒮`$ in terms of renormalized quantities: $`𝒮`$ $`=`$ $`{\displaystyle _0^\beta }d\tau {\displaystyle \underset{k\sigma }{}}\overline{c}_{k\sigma }^{}(_\tau +ϵ_k)c_{k\sigma }^{}`$ $`+{\displaystyle _0^\beta }d\tau {\displaystyle \underset{m}{}}\overline{f}_m^{}(_\tau +ϵ_f)f_m^{}+{\displaystyle \underset{\sigma }{}}\overline{b}_\sigma ^{}_\tau b_\sigma ^{}`$ $`+{\displaystyle _0^\beta }d\tau V\mu ^{\overline{r}}\left[\overline{f}_0b_{}^{}c_{}^{}(\mathrm{𝟎})+\overline{f}_0b_{}^{}c_{}^{}(\mathrm{𝟎})\right]+\mathrm{c}.\mathrm{c}.,`$ $`+{\displaystyle _0^\beta }d\tau \sqrt{2}V\mu ^{\overline{r}}\left[\overline{f}_1b_{}^{}c_{}^{}(\mathrm{𝟎})+\overline{f}_1b_{}^{}c_{}^{}(\mathrm{𝟎})\right]+\mathrm{c}.\mathrm{c}.,`$ $`𝒮_\lambda `$ $`=`$ $`{\displaystyle _0^\beta }d\tau \lambda \left({\displaystyle \underset{m}{}}\overline{f}_m^{}f_m^{}+{\displaystyle \underset{\sigma }{}}\overline{b}_\sigma ^{}b_\sigma ^{}1\right),`$ (48) $`\delta 𝒮`$ $`=`$ $`{\displaystyle _0^\beta }d\tau \delta \lambda _f{\displaystyle \underset{m}{}}\overline{f}_m^{}f_m^{}+\delta \lambda _b{\displaystyle \underset{\sigma }{}}\overline{b}_\sigma ^{}b_\sigma ^{}+\mathrm{}`$ (49) As above, we have introduced a renormalized dimensionful hybridization $`V=v/N_0`$. We did not write the vertex renormalization factor, since it will not appear at this order of the calculation. We also have to introduce the bosonic Green’s function <sup>1</sup><sup>1</sup>1Because of the minus sign involved in this definition, an extra $`(1)`$ factor should be associated to each bosonic lines in the diagrammatics.: $$G_{b0}(i\nu )=<b_\sigma ^{}(i\nu )b_\sigma ^{}(i\nu )>=\frac{1}{i\nu \lambda }.$$ (50) Calculating the self-energy for the $`f_1`$ fermion shown in Fig. 6, we find at zero temperature: $$\mathrm{\Sigma }_f(i\omega )=2N_0V^2\mu ^{2\overline{r}}_0^Ddϵ\frac{ϵ^r}{ϵi\omega }+\delta \lambda _fi\omega (Z_f1)$$ (51) which reduces to (27) when the limit $`\overline{r}0`$ is taken. The calculation for the $`b_{}`$ proceeds along the same line, although the prefactors differ slightly (see Fig. 6). ### B.2 Impurity entropy To evaluate the impurity entropy, we need to expand the physical partition function $`𝒵`$, which is represented diagrammatically in Fig. 7. The free energy can then be obtained through $`F=T\mathrm{ln}𝒵`$, and reads: $`F`$ $`=`$ $`T\mathrm{ln}5+{\displaystyle \frac{12}{5}}v^2\mu ^{2\overline{r}}{\displaystyle _0^D}dϵ{\displaystyle \frac{ϵ^r}{ϵ}}`$ (52) $`{\displaystyle \frac{12}{5}}v^2\mu ^{2\overline{r}}{\displaystyle _0^D}dϵ{\displaystyle \frac{ϵ^r}{ϵ}}\mathrm{tanh}{\displaystyle \frac{ϵ}{2T}}`$ which is the result quoted in Eq. (34) from which the impurity entropy at the ACR fixed point follows. In deriving this result we have used the renormalized values for the counter-terms (29)-(31), which provide the correction in the first line of the above expresion. One can check that they actually drop out when the entropy is calculated, but those corrections are nevertheless crucial for regularizing the forthcoming impurity susceptibility. ### B.3 Impurity susceptibility Computing the impurity susceptibility is quite cumbersome, since many graphs appear at intermediate stages of the calculation, due to the fact that both pseudoparticles now carry a spin index. First, we recall that the quantity which admits a simple diagrammatics is in fact $`<Q>\chi _{\mathrm{imp}}`$ rather than $`\chi _{\mathrm{imp}}`$, where $`<Q>`$ is obtained from $`<Q>=e^{\beta \lambda }𝒵`$ and the above result for the free energy, Eq. (52). Again, we simply quote the collection of graphs that need to be evaluated in Fig. 8 (with the corresponding prefactor), and refer the reader to Ref. fritz1, for more details.
warning/0506/hep-th0506177.html
ar5iv
text
# Contents ## 1 Introduction and Summary Analysis of supersymmetric black holes in string theory have led to many new insights into the classical and quantum aspects of black holes. In particular a rich structure has emerged in the context of half-BPS black holes in $`𝒩=2`$ supersymmetric string theories in four dimensions. One of the important features of these black holes is the attractor mechanism by which the values of the scalar fields at the horizon are determined only by the charges carried by the black hole and are independent of the asymptotic values of the scalar fields. The entropy of these black holes agrees with the microscopic counting of the states of the brane system they describe, not only in the supergravity approximation, but also after the inclusion of higher derivative corrections to the generalized prepotential. More recently it has been shown that the Legendre transform of the black hole entropy with respect to the electric charges is directly related to the generalized prepotential, and this has led to a new conjectured relation between the black hole entropy and topological string partition function. Finally, applying the results for these black holes to the special case of black holes in heterotic string theory with purely electric charges, one finds agreement between black hole entropy and the degeneracy of elementary string states even though the black hole entropy vanishes in the supergravity approximation. All of these results have been derived by making heavy use of supersymmetry. In particular while taking into account the effect of higher derivative terms one includes in the string theory effective action only a special class of terms which can be computed using the partition function of topological string theory. These corrections are controlled by a special function known as the generalized prepotential. While these constitute an important set of terms in the string theory effective acion, they are by no means the only terms, and at present there is no understanding of why these terms should play a special role in the study of black holes. In fact there are counterexamples, involving elementary string states in type II string theory, for which the corrections to the generalized prepotential are not enough to produce the desired result for the black hole entropy. Thus it seems important to study the role of the complete set of higher derivative terms on the near horizon geometry of the black hole. In this paper we study the effect of higher derivative terms on the entropy of extremal black holes in $`D`$ dimensions following the general formalism developed in . We do not make use of supersymmetry directly, but define extremal black holes to be those objects whose near horizon geometry is given by $`AdS_2\times S^{D2}`$.<sup>1</sup><sup>1</sup>1Eventually supersymmetry may play a role in establishing the existence of a solution that interpolates between the near horizon $`AdS_2\times S^{D2}`$ geometry and the asymptotic Minkowski space-time. We also define the entropy of the extremal black hole to be the extremal limit of the entropy of a non-extremal black hole so that we can use the general formula for the entropy given in even though strictly extremal black holes do not have a bifurcate horizon. Our main results may be summarized as follows. 1. Let $`S_{BH}(\stackrel{}{q},\stackrel{}{p})`$ denote the entropy of a D-dimensional extremal black hole, with near horizon geometry $`AdS_2\times S^{D2}`$, as a function of electric charges $`\{q_i\}`$ associated with one form gauge fields and magnetic charges $`\{p_a\}`$ associated with $`(D3)`$ form gauge fields. We choose a coordinate system in which the $`AdS_2`$ part of the metric is proportional to $`r^2dt^2+dr^2/r^2`$. Then the Legendre transform of $`S_{BH}(\stackrel{}{q},\stackrel{}{p})/2\pi `$ with respect to the variables $`q_i`$ is equal to the the integral of the Lagrangian density over the $`(D2)`$ dimensional sphere $`S^{D2}`$ enclosing the black hole. The variable conjugate to $`q_i`$ represents the radial electric field $`e_i`$ at the horizon associated with the $`i`$-th gauge field. 2. Consider a general $`AdS_2\times S^{D2}`$ background parametrized by the sizes of $`AdS_2`$ and $`S^{D2}`$, the electric and magnetic fields and the values of various scalar fields. We define an entropy function by integrating the Lagrangian density evaluated for this background over $`S^{D2}`$, taking the Legendre transform of this integral with respect to the parameters $`e_i`$ labelling the electric fields and multiplying the result by $`2\pi `$. The result is a function of the values $`u_s`$ of the scalar fields, the sizes $`v_1`$ and $`v_2`$ of $`AdS_2`$ and $`S^{D2}`$, the electric charges $`q_i`$ conjugate to the variables $`e_i`$, and the magnetic charges $`p_a`$ labelling the background magnetic fields. We show that for given $`\stackrel{}{q}`$ and $`\stackrel{}{p}`$, the values $`u_s`$ of the scalar fields as well as the sizes $`v_1`$ and $`v_2`$ of $`AdS_2`$ and $`S^{D2}`$ are determined by extremizing the entropy function with respect to the variables $`u_i`$, $`v_1`$ and $`v_2`$. Furthermore the entropy itself is given by the value of the entropy function at the horizon. 3. For extremal black hole solutions without Ramond-Ramond (RR) charges in tree level string theory the Lagrangian density at the horizon vanishes due to the dilaton field equation. In this case the entropy of the black hole is given simply by $`2\pi `$ times the product of the electric field at the horizon and the electric charge of the black hole. These results rely on the assumption that the Lagrangian density can be expressed in terms of gauge invariant field strengths and does not involve the gauge fields explicitly. Thus if Chern-Simons terms are present we either need to remove them by going to the dual field variables, or if that is not possible, consider black hole solutions which are not affected by these Chern-Simons terms. ## 2 Entropy of Extremal Black Holes We begin by considering a four dimensional theory of gravity coupled to a set of abelian gauge fields $`A_\mu ^{(i)}`$ and neutral scalar fields $`\{\varphi _s\}`$. Suppose $`\sqrt{detg}`$ is the lagrangian density, expressed as a function of the metric $`g_{\mu \nu }`$, the scalar fields $`\{\varphi _s\}`$, the gauge field strengths $`F_{\mu \nu }^{(i)}`$, and covariant derivatives of these fields. We consider a spherically symmetric extremal black hole solution with near horizon geometry $`AdS_2\times S^2`$. The most general field configuration, consistent with the $`SO(2,1)\times SO(3)`$ symmetry of $`AdS_2\times S^2`$, is of the form: $`ds^2g_{\mu \nu }dx^\mu dx^\nu =v_1\left(r^2dt^2+{\displaystyle \frac{dr^2}{r^2}}\right)+v_2\left(d\theta ^2+\mathrm{sin}^2\theta d\varphi ^2\right)`$ $`\varphi _s=u_s`$ $`F_{rt}^{(i)}=e_i,F_{\theta \varphi }^{(i)}={\displaystyle \frac{p_i}{4\pi }}\mathrm{sin}\theta ,`$ (2.1) where $`v_1`$, $`v_2`$, $`\{u_s\}`$, $`\{e_i\}`$ and $`\{p_i\}`$ are constants. For this background the nonvanishing components of the Riemann tensor are:<sup>2</sup><sup>2</sup>2In our convention $`R_{\nu \rho \sigma }^\mu =_\rho \mathrm{\Gamma }_{\nu \sigma }^\mu _\sigma \mathrm{\Gamma }_{\nu \rho }^\mu +\mathrm{\Gamma }_{\tau \rho }^\mu \mathrm{\Gamma }_{\nu \sigma }^\tau \mathrm{\Gamma }_{\tau \sigma }^\mu \mathrm{\Gamma }_{\nu \rho }^\tau `$ where $`\mathrm{\Gamma }_{\nu \rho }^\mu `$ is the Christoffel symbol. $`R_{\alpha \beta \gamma \delta }`$ $`=`$ $`v_1^1(g_{\alpha \gamma }g_{\beta \delta }g_{\alpha \delta }g_{\beta \gamma }),\alpha ,\beta ,\gamma ,\delta =r,t,`$ $`R_{mnpq}`$ $`=`$ $`v_2^1(g_{mp}g_{nq}g_{mq}g_{np}),m,n,p,q=\theta ,\varphi .`$ (2.2) It follows from the general form of the background that the covariant derivatives of the scalar fields $`\varphi _s`$, the gauge field strengths $`F_{\mu \nu }^{(i)}`$ and the Riemann tensor $`R_{\mu \nu \rho \sigma }`$ all vanish for the near horizon geometry. By the general symmetry consideration it follows that the contribution to the equation of motion from any term in the action that involves covariant derivatives of the gauge field strengths, scalars or the Riemann tensor vanish identically for this background and we can restrict our attention to only those terms which do not involve covariant derivatives of these fields.<sup>3</sup><sup>3</sup>3We are assuming that all terms in the action depend explicitly only on the gauge field strengths and not on gauge fields. This condition is violated for example in string theory by Chern-Simons type coupling of the gauge fields to three form field strengths. However, as is well known, we can get rid of such terms by dualizing the two form field to a scalar axion $`a`$. This field couples to the gauge fields only through field strengths. If we encounter a theory where it is impossible to carry this out for all fields, our analysis will still be valid if these additional terms do not affect the equation of motion and the entropy for the specific black hole solution under study. Let us denote by $`f(\stackrel{}{u},\stackrel{}{v},\stackrel{}{e},\stackrel{}{p})`$ the Lagrangian density $`\sqrt{detg}`$ evaluated for the near horizon geometry (LABEL:e1) and integrated over the angular coordinates: $$f(\stackrel{}{u},\stackrel{}{v},\stackrel{}{e},\stackrel{}{p})=𝑑\theta 𝑑\varphi \sqrt{detg}.$$ (2.3) The scalar and the metric field equations in the near horizon geometry correspond to extremizing $`f`$ with respect to the variables $`\stackrel{}{u}`$ and $`\stackrel{}{v}`$: $$\frac{f}{u_s}=0,\frac{f}{v_i}=0.$$ (2.4) On the other hand the non-trivial components of the gauge field equations and the Bianchi identities take the form: $$_r\left(\frac{\sqrt{detg}}{F_{rt}^{(i)}}\right)=0,_rF_{\theta \varphi }^{(i)}=0.$$ (2.5) Both sets of equations in (2.5) are automatically satisfied by the background (LABEL:e1), with the constants of integration having the interpretation as electric and magnetic charges of the black hole. From this it follows that the constants $`p_i`$ appearing in (LABEL:e1) correspond to magnetic charges of the black hole, and $$\frac{f}{e_i}=q_i$$ (2.6) where $`q_i`$ denote the electric charges carried by the black hole. For fixed $`\stackrel{}{p}`$ and $`\stackrel{}{q}`$, (2.4) and (2.6) give a set of equations which are equal in number to the number of unknowns $`\stackrel{}{u}`$, $`\stackrel{}{v}`$ and $`\stackrel{}{e}`$. In a generic case we may be able to solve these equations completely to determine the background in terms of only the electric and the magnetic charges $`\stackrel{}{q}`$ and $`\stackrel{}{p}`$. <sup>4</sup><sup>4</sup>4We should note however that the situation in string theory is not completely generic. For example in $`𝒩=2`$ supersymmetric string theories there is no coupling of the hypermultiplet scalars to the vector multiplet fields or the curvature tensor to lowest order in $`\alpha ^{}`$, and hence in this approximation the function $`f`$ does not depend on the hypermultiplet scalars. Thus the equations (2.4), (2.6) do not fix the values of the hypermultiplet scalars in this approximation. This is consistent with the attractor mechanism for supersymmetric background which says that the near horizon configuration of a black hole depends only on the electric and magnetic charges carried by the black hole and not on the asymptotic values of these scalar fields. We shall return to a more detailed discussion of this mechanism in section 3. Let us now turn to the analysis of the entropy associated with this black hole. A general formula for the entropy in the presence of higher derivative terms has been given in . The formula simplifies enormously here since the covariant derivatives of all the tensors vanish, and we get a simple formula: $$S_{BH}=8\pi \frac{}{R_{rtrt}}g_{rr}g_{tt}A_H,$$ (2.7) where $`A_H`$ is the area of the event horizon and $`\frac{}{R_{\mu \nu \rho \sigma }}`$ is defined through the equation $$\delta =\frac{}{R_{\mu \nu \rho \sigma }}\delta R_{\mu \nu \rho \sigma }.$$ (2.8) In computing $`\delta `$ we can ignore all terms in $``$ which involve covariant derivatives of the Riemann tensor, and treat the components of the Riemann tensor as independent variables. In order to simplify this formula let us denote by $`f_\lambda (\stackrel{}{u},\stackrel{}{v},\stackrel{}{e},\stackrel{}{p})`$ an expression similar to the right hand side of (2.3) except that each factor of $`R_{rtrt}`$ in the expression of $``$ is multiplied by a factor of $`\lambda `$. Then we have the relation: $$\frac{f_\lambda (\stackrel{}{u},\stackrel{}{v},\stackrel{}{e},\stackrel{}{p})}{\lambda }|_{\lambda =1}=𝑑\theta 𝑑\varphi \sqrt{detg}R_{\alpha \beta \gamma \delta }\frac{}{R_{\alpha \beta \gamma \delta }},$$ (2.9) where the repeated indices $`\alpha ,\beta ,\gamma ,\delta `$ are summed over the coordinates $`r`$ and $`t`$. Now since by symmetry consideration $`(/R_{\alpha \beta \gamma \delta })`$ is proportional to $`(g^{\alpha \gamma }g^{\beta \delta }g^{\alpha \delta }g^{\beta \gamma })`$, we have $$\frac{}{R_{\alpha \beta \gamma \delta }}=v_1^2(g^{\alpha \gamma }g^{\beta \delta }g^{\alpha \delta }g^{\beta \gamma })\frac{}{R_{rtrt}}.$$ (2.10) The constant of proportionality has been fixed by taking $`(\alpha \beta \gamma \delta )=(rtrt)`$. Using (2) and (2.10) we can rewrite (2.9) as $$\frac{}{R_{rtrt}}A_H=\frac{1}{4}v_1^2\frac{f_\lambda (\stackrel{}{u},\stackrel{}{v},\stackrel{}{e},\stackrel{}{p})}{\lambda }|_{\lambda =1}.$$ (2.11) Substituting this into (2.7) gives $$S_{BH}=2\pi \frac{f_\lambda (\stackrel{}{u},\stackrel{}{v},\stackrel{}{e},\stackrel{}{p})}{\lambda }|_{\lambda =1}.$$ (2.12) We shall now reexpress the right hand side of (2.12) in terms of derivatives of $`f`$ with respect to the variables $`\stackrel{}{u}`$, $`\stackrel{}{v}`$, $`\stackrel{}{e}`$ and $`\stackrel{}{p}`$. Since the expression for $``$ is invariant under reparametrization of the $`r,t`$ coordinates, every factor of $`R_{rtrt}`$ in the expression for $`f_\lambda `$ must appear in the combination $`\lambda g^{rr}g^{tt}R_{rtrt}=\lambda v_1^1`$, every factor of $`F_{rt}^{(i)}`$ must appear in the combination $`\sqrt{g^{rr}g^{tt}}F_{rt}^{(i)}=e_iv_1^1`$, and every factor of $`F_{\theta \varphi }^{(i)}=e_i`$ and $`\varphi _s=u_s`$ must appear without any accompanying power of $`v_1`$. The contribution from all terms which involve covariant derivatives of $`F_{\mu \nu }^{(i)}`$, $`R_{\mu \nu \rho \sigma }`$ or $`\varphi _s`$ vanish; hence there is no further factor of $`v_1`$ coming from contraction of the metric with these derivative operators. The only other $`v_1`$ dependence of $`f_\lambda (\stackrel{}{u},\stackrel{}{v},\stackrel{}{e},\stackrel{}{p})`$ is through the overall multiplicative factor of $`\sqrt{detg}v_1`$. Thus $`f_\lambda (\stackrel{}{u},\stackrel{}{v},\stackrel{}{e},\stackrel{}{p})`$ must be of the form $`v_1g(\stackrel{}{u},v_2,\stackrel{}{p},\lambda v_1^1,\stackrel{}{e}v_1^1)`$ for some function $`g`$, and we have $$\lambda \frac{f_\lambda (\stackrel{}{u},\stackrel{}{v},\stackrel{}{e},\stackrel{}{p})}{\lambda }+v_1\frac{f_\lambda (\stackrel{}{u},\stackrel{}{v},\stackrel{}{e},\stackrel{}{p})}{v_1}+e_i\frac{f_\lambda (\stackrel{}{u},\stackrel{}{v},\stackrel{}{e},\stackrel{}{p})}{e_i}f_\lambda (\stackrel{}{u},\stackrel{}{v},\stackrel{}{e},\stackrel{}{p})=0.$$ (2.13) Setting $`\lambda =1`$ in (2.13), using the equation of motion of $`v_1`$ as given in (2.4), and substituting the result into eq.(2.12) we get $$S_{BH}=2\pi \left(e_i\frac{f}{e_i}f\right).$$ (2.14) This together with (2.6) shows that $`S_{BH}(\stackrel{}{q},\stackrel{}{p})/2\pi `$ may be regarded as the Legendre transform of the function $`f(\stackrel{}{u},\stackrel{}{v},\stackrel{}{e},\stackrel{}{p})`$ with respect to the variables $`e_i`$ after eliminating $`\stackrel{}{u}`$ and $`\stackrel{}{v}`$ through their equations of motion (2.4). The analysis can be easily generalized to higher dimensional theories as follows. In $`D`$ space-time dimensions we consider an extremal black hole solution with near horizon geometry $`AdS_2\times S^{D2}`$. The relevant fields which can take non-trivial expectation value near the horizon are scalars $`\{\varphi _s\}`$, metric $`g_{\mu \nu }`$, gauge fields $`A_\mu ^{(i)}`$ and $`(D3)`$-form gauge fields $`B_{\mu _1\mathrm{}\mu _{D3}}^{(a)}`$. If $`H_{\mu _1\mathrm{}\mu _{D2}}^{(a)}`$ denote the field strength associated with the $`B`$ field, then the general background consistent with the $`SO(2,1)\times SO(D1)`$ symmetry of the background geometry is of the form: $`ds^2g_{\mu \nu }dx^\mu dx^\nu =v_1\left(r^2dt^2+{\displaystyle \frac{dr^2}{r^2}}\right)+v_2d\mathrm{\Omega }_{D2}^2`$ $`\varphi _s=u_s`$ $`F_{rt}^{(i)}=e_i,H_{l_1\mathrm{}l_{D2}}^{(a)}=p_aϵ_{l_1\mathrm{}l_{D2}}\sqrt{deth^{(D2)}}/\mathrm{\Omega }_{D2}.`$ (2.15) where $`d\mathrm{\Omega }_{D2}=h_{ll^{}}^{(D2)}dx^ldx^l^{}`$ denotes the line element on the unit $`(D2)`$-sphere, $`\mathrm{\Omega }_{D2}`$ denotes the area of the unit $`(D2)`$-sphere, $`x^{l_i}`$ with $`2l_i(D1)`$ are coordinates along this sphere and $`ϵ`$ denotes the totally anti-symmetric symbol with $`ϵ_{2\mathrm{}(D1)}=1`$. We now define $$f(\stackrel{}{u},\stackrel{}{v},\stackrel{}{e},\stackrel{}{p})=𝑑x^2\mathrm{}𝑑x^{D1}\sqrt{detg},$$ (2.16) as in (2.3). Analysis identical to that for $`D=4`$ now tells us that the constants $`p_a`$ represent magnetic type charges carried by the black hole, and the equations which determine the values of $`\stackrel{}{u}`$, $`\stackrel{}{v}`$ and $`\stackrel{}{e}`$ are $$\frac{f}{u_s}=0,\frac{f}{v_i}=0,\frac{f}{e_i}=q_i,$$ (2.17) where $`q_i`$ denote the electric charges carried by the black hole. Also using (2.7) which is valid in any dimension, we can show that the entropy of the black hole is given by $`2\pi `$ times the Legendre transform of $`f`$: $$S_{BH}=2\pi \left(e_i\frac{f}{e_i}f\right).$$ (2.18) as in (2.14). At string tree level, and in the absence of Ramond-Ramond background fields (which includes all black holes in heterotic string theory) the Lagrangian density at the horizon and hence the function $`f`$ vanishes due to the dilaton field equation. Thus eqs.(2.17), (2.18) give: $$S_{BH}=2\pi q_ie_i.$$ (2.19) In other words the entropy of these black holes is given by $`2\pi `$ times the product of the electric charge and the electric field at the horizon. It will be interesting to see if this quantity admits a simple interpretation in the world-sheet conformal field theory that describes this background. ## 3 Attractor Mechanism and the Entropy Function We can now reformulate the attractor mechanism in a more suggestive manner. Let us define $$F(\stackrel{}{u},\stackrel{}{v},\stackrel{}{q},\stackrel{}{p})=2\pi \left(e_i\frac{f(\stackrel{}{u},\stackrel{}{v},\stackrel{}{e},\stackrel{}{p})}{e_i}f(\stackrel{}{u},\stackrel{}{v},\stackrel{}{e},\stackrel{}{p})\right),$$ (3.1) with $`e_i`$ determined by the equation: $$\frac{f(\stackrel{}{u},\stackrel{}{v},\stackrel{}{e},\stackrel{}{p})}{e_i}=q_i.$$ (3.2) In that case it follows from (2.17) that the values of $`\stackrel{}{u}`$ and $`\stackrel{}{v}`$ at the horizon are determined by extremizing the function $`F(\stackrel{}{u},\stackrel{}{v},\stackrel{}{q},\stackrel{}{p})`$ with respect to $`\stackrel{}{u}`$ and $`\stackrel{}{v}`$: $$\frac{F(\stackrel{}{u},\stackrel{}{v},\stackrel{}{q},\stackrel{}{p})}{u_s}=0,\frac{F(\stackrel{}{u},\stackrel{}{v},\stackrel{}{q},\stackrel{}{p})}{v_i}=0.$$ (3.3) Furthermore, eq.(2.18) shows that the black hole entropy $`S_{BH}`$ is given by the value of the function $`F`$ at this extremum: $$S_{BH}(\stackrel{}{q},\stackrel{}{p})=F(\stackrel{}{u},\stackrel{}{v},\stackrel{}{q},\stackrel{}{p}),$$ (3.4) with $`\stackrel{}{u}`$, $`\stackrel{}{v}`$ given by eq.(3.3). This suggests that we call $`F(\stackrel{}{u},\stackrel{}{v},\stackrel{}{q},\stackrel{}{p})`$ the entropy function. Finally, the near horizon electric field $`e_i`$ are given by $$e_i=\frac{1}{2\pi }\frac{F}{q_i}.$$ (3.5) ## 4 Relation to Earlier Results We are now in a position to discuss the relation between our results and the observation of that the Legendre transform of the entropy of a black hole in $`𝒩=2`$ supersymmetric string theory is given by the imaginary part of the generalized prepotential of the theory. In the argument of the prepotential the real parts of the complex vector multiplet scalar fields are replaced, up to a constant of proportionality, by the magnetic charges of the black hole, whereas the imaginary parts of these scalar fields are replaced by the variables conjugate to the electric charges of the black hole. This result follows from our results together with the following observations (see e.g. ): 1. For the near horizon configuration of the black hole in $`𝒩=2`$ supersymmetric string theory, all terms in the Lagrangian density vanish, except for a single term proportional to the imaginary part of the generalized prepotential . 2. For the near horizon geometry the real parts of the vector multiplet scalar fields are proportional to the magnetic field at the horizon whereas the imaginary parts of these scalar fields are proportional to the electric field at the horizon. A little algebra shows that all the normalization factors also work out correctly and we can reproduce the abovementioned observation of from our results. Acknowledgement: I wish to thank Rajesh Gopakumar for his comments on the manuscript. I also wish to thank the members of the Center for Theoretical Physics at MIT and DAMTP at Cambridge University for discussion during various stages of this work. The work was supported in part by the Jane Morningstar visiting professorship at the Center for Theoretical Physics at MIT.
warning/0506/cond-mat0506698.html
ar5iv
text
# Computation of dynamical correlation functions of Heisenberg chains: the gapless anisotropic regime ## I Introduction In recent years, enormous progress has been made in the quest to overcome an important and long-standing limitation of the Bethe Ansatz BetheZP71 framework in the theory of integrable models: the inability to compute correlation functions. Most efforts have been focused on the Heisenberg spin chain JimboBOOK ; KitanineNPB554 ; KitanineNPB567 ; Boosh0412191 , although similiar sets of results could in principle be obtained for other integrable models formulated via the Algebraic Bethe Ansatz KorepinBOOK . In the particular case of the $`XXZ`$ model, matrix elements of any local operator between two Bethe states can now be written as determinants of matrices, whose elements are known analytic functions of the rapidities of the eigenstates involved KitanineNPB554 . Used in conjunction with formulas for eigenstate norms KorepinCMP86 , this yields exact expressions for form factors on the lattice, thereby permitting in principle the computation of dynamical correlation functions. In this paper, we concentrate on the anisotropic Heisenberg spin-1/2 chain in a magnetic field, with Hamiltonian $`H=J{\displaystyle \underset{j=1}{\overset{N}{}}}\left[S_j^xS_{j+1}^x+S_j^yS_{j+1}^y+\mathrm{\Delta }\left(S_j^zS_{j+1}^z{\displaystyle \frac{1}{4}}\right)hS_j^z\right]`$ (1) and periodic boundary conditions. The anisotropy parameter $`\mathrm{\Delta }`$ can take on any real value without destroying the applicability of the Bethe Ansatz or of the present method. We will however consider here only the quantum critical regime $`1<\mathrm{\Delta }1`$. Our interest lies in space- and time-dependent spin-spin correlation functions. Continuing recent work by two of us Caux0502365 , we numerically obtain the dynamical spin-spin structure factor, defined as the Fourier transform of the connected spin-spin correlation function $`S^{a\overline{a}}(q,\omega )={\displaystyle \frac{1}{N}}{\displaystyle \underset{j,j^{}=1}{\overset{N}{}}}e^{iq(jj^{})}{\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑te^{i\omega t}S_j^a(t)S_j^{}^{\overline{a}}(0)_c`$ (2) where $`a=z,\pm `$. This quantity is of fundamental importance in many settings, and is used for example in the quantitative description of inelastic neutron scattering data for quasi-one-dimensional systems KenzelmannPRB65 ; StonePRL91 ; ZaliznyakPRL93 ; LakeNature4 to more theoretical questions relating to quantum entanglement JinJPA69 . In any case, the main interest is based on the fact that the Heisenberg spin chain is a strongly coupled quantum system with fractionalized (spinon-like and higher) excitations, for which it is possible to obtain nonperturbative information. The use of Bethe Ansatz-based methods to obtain quantitative results on correlation functions, in view of the long history of integrability, is rather recent. In the case of zero field, much progress has been possible for the model in the gapped phase JimboBOOK , and for the isotropic limit, for which both two-spinon BougourziPRB54 ; KarbachPRB55 and four-spinon AbadaNPB497 contributions were obtained analytically. The nonzero field problem remains intractable with this approach. However, summations over intermediate states can be performed numerically once the analytical expressions for the form factors are known. Using this strategy, the spin-spin correlations (both longitudinal and transverse) for the isotropic Heisenberg model in a field were studied at fixed momentum $`q`$ in \[BiegelEPL59, \], and the transverse ones for $`XXZ`$ at zero field and at $`q=\pi `$ in \[BiegelJPA36, \]. Two-particle contributions to the longitudinal structure factor for the $`XXZ`$ chain in a field at $`q=\pi /2`$ were also studied in \[SatoJPSJ73, \]. Multiparticle contributions and extension to the full Brillouin zone was done in \[Caux0502365, \]. In these, it is observed that summing over a relatively small subset of intermediate states is sufficient to obtain good precision, as demonstrated by using sum rules. The present paper has three main further objectives: to generalize the form factor determinant formulae to intermediate states containing complex rapidities (in the form of strings) present in the Bethe Ansatz for the $`XXZ`$ chain, to study the general field dependence of the correlation functions, and to obtain the real space- and time-dependent correlators. The plan of the paper is as follows. We begin in section II by recalling a few basic facts concerning the Bethe Ansatz for the $`XXZ`$ model, and the lattice form factors for single spin operators. In section III, we discuss string solutions to the Bethe equations for excited states, and obtain modified determinant representations for the form factors of local spin operators involving such states. Section IV contains the results of our numerical evaluations of the form factor sums for dynamical spin-spin correlation functions for two example choices of anisotropy. In section V, we use our results to compute the space- and time-dependent correlators. This allows us in particular to obtain equal-time correlators for any lattice separation. Comparison is made with known exact results in zero field for small lattice separation. We also compare our results to predictions from conformal field theory for the asymptotics in zero field. Our conclusions and outlook are collected in section VI. ## II Setup In this section, we briefly review all elements necessary for the computation of the spin-spin correlation functions of the anisotropic Heisenberg model. The exact solution through the Bethe Ansatz for model (1) is well-known (see \[KorepinBOOK, ; TakahashiBOOK, \] and references therein). The reference state is taken to be the state with all spins up, $`|0=_{i=1}^N|_i`$. Since the total magnetization commutes with the Hamiltonian, the Hilbert space separates into subspaces of fixed magnetization, determined from the number of reversed spins $`M`$. We take the number of sites $`N`$ to be even, and $`2MN`$, the other sector being accessible through a change in the reference state. Eigenstates in each subspace are completely characterized for $`2MN`$ by a set of rapidities $`\{\lambda _j\}`$, $`j=1,\mathrm{},M`$, solution to the Bethe equations $`\left[{\displaystyle \frac{\mathrm{sinh}(\lambda _j+i\zeta /2)}{\mathrm{sinh}(\lambda _ji\zeta /2)}}\right]^N={\displaystyle \underset{kj}{\overset{M}{}}}{\displaystyle \frac{\mathrm{sinh}(\lambda _j\lambda _k+i\zeta )}{\mathrm{sinh}(\lambda _j\lambda _ki\zeta )}},j=1,\mathrm{},M`$ (3) where $`\mathrm{\Delta }=\mathrm{cos}\zeta `$. In view of the periodicity of the $`\mathrm{sinh}`$ function in the complex plane, we can restrict the possible values that the rapidities can take to the strip $`\pi /2<\text{Im}\lambda \pi /2`$, or alternately define an extended zone scheme in which $`\lambda `$ and $`\lambda +i\pi `$ are identified. A more practical version of the Bethe equations is obtained by writing them in logarithmic form, $`\text{atan}[{\displaystyle \frac{\mathrm{tanh}(\lambda _j)}{\mathrm{tan}(\zeta /2)}}]{\displaystyle \frac{1}{N}}{\displaystyle \underset{k=1}{\overset{M}{}}}\text{atan}[{\displaystyle \frac{\mathrm{tanh}(\lambda _j\lambda _k)}{\mathrm{tan}\zeta }}]=\pi {\displaystyle \frac{I_j}{N}}.`$ (4) Here, $`I_j`$ are distinct half-integers which can be viewed as quantum numbers: each choice of a set $`\{I_j\}`$, $`j=1,\mathrm{},M`$ (with $`I_j`$ defined mod$`(N)`$) uniquely specifies a set of rapidities, and therefore an eigenstate. The energy of a state is given as a function of the rapidities by $`E=J{\displaystyle \underset{j=1}{\overset{M}{}}}{\displaystyle \frac{\mathrm{sin}^2\zeta }{\mathrm{cosh}2\lambda _j\mathrm{cos}\zeta }}h({\displaystyle \frac{N}{2}}M),`$ (5) whereas the momentum has a simple representation in terms of the quantum numbers, $`q={\displaystyle \underset{j=1}{\overset{M}{}}}i\mathrm{ln}\left[{\displaystyle \frac{\mathrm{sinh}(\lambda _j+i\zeta /2)}{\mathrm{sinh}(\lambda _ji\zeta /2)}}\right]=\pi M+{\displaystyle \frac{2\pi }{N}}{\displaystyle \underset{j=1}{\overset{M}{}}}I_j\text{mod}2\pi .`$ (6) The ground state is given by $`I_j^0=\frac{M+1}{2}+j`$, $`j=1,\mathrm{},M`$, and all excited states are in principle obtained from the different choices of sets $`\{I_j\}`$. To study dynamics, some ingredients have to be added to the Bethe Ansatz: the matrix elements of spin operators between eigenstates (form factors). In terms of form factors for the Fourier-transformed spin operators $`S_q^a=\frac{1}{\sqrt{N}}_{j=1}^Ne^{iqj}S_j^a`$, the structure factor (2) can be written as a sum $`S^{a\overline{a}}(q,\omega )=2\pi {\displaystyle \underset{\alpha }{}}|GS|S_q^a|\alpha |^2\delta (\omega \omega _\alpha )`$ (7) over the whole set of contributing intermediate eigenstates $`|\alpha `$. For the longitudinal structure factor, this is the set of all states with the same number of overturned spins as the ground state in the chosen magnetization subsector, excluding the ground state itself (connected correlator). For the transverse structure factor $`S^+`$, it is the set of all states with one less overturned spin. $`\omega _\alpha `$ is the energy difference of state $`|\alpha `$ with the ground state. Each term in (7) can be obtained KitanineNPB554 as a product of determinants of specific matrices, which are fully determined for given bra and ket eigenstates by a knowledge of the corresponding sets of rapidities. The form factor between two eigenstates (with $`M`$ reversed spins, and rapidities $`\{\mu \}`$, $`\{\lambda \}`$) for the $`S^z`$ operator is given by KitanineNPB554 $`|\{\mu \}|S_q^z|\{\lambda \}|^2={\displaystyle \frac{N}{4}}\delta _{q,q_{\{\lambda \}}q_{\{\mu \}}}{\displaystyle \underset{j=1}{\overset{M}{}}}|{\displaystyle \frac{\mathrm{sinh}(\mu _ji\zeta /2)}{\mathrm{sinh}(\lambda _ji\zeta /2)}}|^2{\displaystyle \underset{j>k=1}{\overset{M}{}}}|\mathrm{sinh}^2(\mu _j\mu _k)+\mathrm{sin}^2\zeta |^1\times `$ $`\times {\displaystyle \underset{j>k=1}{\overset{M}{}}}|\mathrm{sinh}^2(\lambda _j\lambda _k)+\mathrm{sin}^2\zeta |^1{\displaystyle \frac{\left|det[𝐇(\{\mu \},\{\lambda \})2𝐏(\{\mu \},\{\lambda \})]\right|^2}{\left|det𝚽(\{\mu \})det𝚽(\{\lambda \})\right|}}`$ (8) where the matrices $`H`$ and $`P`$ are defined as $`𝐇_{ab}(\{\mu \},\{\lambda \})={\displaystyle \frac{1}{\mathrm{sinh}(\mu _a\lambda _b)}}\left[{\displaystyle \underset{ja}{\overset{M}{}}}\mathrm{sinh}(\mu _j\lambda _bi\zeta )\left[{\displaystyle \frac{\mathrm{sinh}(\lambda _b+i\zeta /2)}{\mathrm{sinh}(\lambda _bi\zeta /2)}}\right]^N{\displaystyle \underset{ja}{\overset{M}{}}}\mathrm{sinh}(\mu _j\lambda _b+i\zeta )\right],`$ (9) $`𝐏_{ab}(\{\mu \},\{\lambda \})={\displaystyle \frac{1}{\mathrm{sinh}^2\mu _a+\mathrm{sin}^2\zeta /2}}{\displaystyle \underset{k=1}{\overset{M}{}}}\mathrm{sinh}(\lambda _k\lambda _bi\zeta ),`$ (10) Similarly, the form factor for the $`S^{}`$ operator between eigenstates with $`M`$ and $`M1`$ reversed spins (respectively having sets of rapidities given by $`\{\mu \}`$, $`\{\lambda \}`$) is KitanineNPB554 $`|\{\mu \}|S_q^{}|\{\lambda \}|^2=N\delta _{q,q_{\{\lambda \}}q_{\{\mu \}}}|\mathrm{sin}\zeta |{\displaystyle \frac{_{j=1}^M\left|\mathrm{sinh}(\mu _ji\zeta /2)\right|^2}{_{j=1}^{M1}\left|\mathrm{sinh}(\lambda _ji\zeta /2)\right|^2}}{\displaystyle \underset{j>k=1}{\overset{M}{}}}|\mathrm{sinh}^2(\mu _j\mu _k)+\mathrm{sin}^2\zeta |^1\times `$ $`\times {\displaystyle \underset{j>k=1}{\overset{M1}{}}}|\mathrm{sinh}^2(\lambda _j\lambda _k)+\mathrm{sin}^2\zeta |^1{\displaystyle \frac{\left|det𝐇^{}(\{\mu \},\{\lambda \})\right|^2}{\left|det𝚽(\{\mu \})det𝚽(\{\lambda \})\right|}}`$ (11) in which the $`H^{}`$ matrix is defined as $`𝐇_{ab}^{}(\{\mu \},\{\lambda \})={\displaystyle \frac{1}{\mathrm{sinh}(\mu _a\lambda _b)}}\left[{\displaystyle \underset{ja}{\overset{M}{}}}\mathrm{sinh}(\mu _j\lambda _bi\zeta )\left[{\displaystyle \frac{\mathrm{sinh}(\lambda _b+i\zeta /2)}{\mathrm{sinh}(\lambda _bi\zeta /2)}}\right]^N{\displaystyle \underset{ja}{\overset{M}{}}}\mathrm{sinh}(\mu _j\lambda _b+i\zeta )\right],b<M,`$ $`𝐇_{aM}^{}(\{\mu \},\{\lambda \})={\displaystyle \frac{1}{\mathrm{sinh}^2\mu _a+\mathrm{sin}^2\zeta /2}}.`$ (12) The norm of the eigenstates appears in the above expressions in the form of the determinant of the Gaudin matrix GaudinBOOK ; KorepinCMP86 $`𝚽_{ab}(\{\lambda \})=\delta _{ab}\left[N{\displaystyle \frac{\mathrm{sin}\zeta }{\mathrm{sinh}^2\lambda _a+\mathrm{sin}^2\zeta /2}}{\displaystyle \underset{ka}{}}{\displaystyle \frac{\mathrm{sin}2\zeta }{\mathrm{sinh}^2(\lambda _a\lambda _k)+\mathrm{sin}^2\zeta }}\right]+(1\delta _{ab}){\displaystyle \frac{\mathrm{sin}2\zeta }{\mathrm{sinh}^2(\lambda _a\lambda _b)+\mathrm{sin}^2\zeta }}.`$ (13) ## III Strings Although the ground state of the $`XXZ`$ chain is characterized by real rapidities, this is not true for all excited states. It has been known ever since the original work of Bethe BetheZP71 that there exist sets of complex rapidities satisfying the Bethe equations. Typically, complex rapidities come in conjugate pairs or more elaborate structures involving higher numbers of elements. These complex rapidity groupings represent bound states of overturned spins, with spatial extent inversely related to the imaginary parts. At low densities (so at high magnetic fields), most solutions take the form of so-called strings, in which a number of rapidities share a real center but have regularly-spaced imaginary parts. For the $`XXZ`$ model, a classification of the possible types of string solutions has been conjectured in \[TakahashiPTP46, \], and further justified in \[HidaPLA84, ; FowlerPRB24, \]. This classification depends sensitively on the value of the anisotropy parameter $`\mathrm{\Delta }`$, and string state counting is traditionally used to argue for the completeness of the string state basis both for the isotropic magnet TakahashiPTP46\_1 ; GaudinBOOK as for the anisotropic one KirillovJPA30 . Our principal aim in this section is to adapt the determinant formulae for form factors and state norms to the case where one of the eigenstates involved contains string solutions to the Bethe equations. This is a necessary procedure, as both the Bethe equations and the determinant formulae for form factors suffer from degenerate limits in the presence of strings, which have to be analytically dealt with by hand. Let us however begin by reviewing the string structure, and setting out our conventions and notations. Following Takahashi and Suzuki TakahashiPTP46 , we take $`\zeta /\pi `$ to be a real number between $`0`$ and $`1`$, which is expressed as a continued fraction of real positive integers as $`{\displaystyle \frac{\zeta }{\pi }}={\displaystyle \frac{1|}{|\nu _1}}+{\displaystyle \frac{1|}{|\nu _2}}+\mathrm{}{\displaystyle \frac{1|}{|\nu _l}},\nu _1,\mathrm{},\nu _{l1}1,\nu _l2.`$ (14) For large $`N`$, rapidities congregate to form strings centered either on the real line (for positive parity strings) or on the axis $`i\pi /2`$ (for negative parity strings), $`\lambda _\alpha ^{n_j,a}=\lambda _\alpha ^{n_j}+i{\displaystyle \frac{\zeta }{2}}(n_j+12a)+i{\displaystyle \frac{\pi }{4}}(1v_j)+i\delta _\alpha ^{n_j,a},a=1,\mathrm{},n_j`$ (15) where the allowable lengths $`n_j`$ and parities $`v_j=\pm 1`$ are to be determined. In a string configuration, the parameters $`\delta _\alpha ^{n_j,a}`$ are exponentially suppressed with system size. The classification of allowable string types in the thermodynamic limit proceeds according to the following algorithm. First, the positive integer series $`y_1,y_0,y_1,\mathrm{},y_l`$ and $`m_0,m_1,\mathrm{},m_l`$ are defined as $`y_1=0,y_0=1,y_1=\nu _1,\text{and}y_i=y_{i2}+\nu _iy_{i1},i=2,\mathrm{},l,m_0=0,m_i={\displaystyle \underset{k=1}{\overset{i}{}}}\nu _k.`$ (16) Lengths and parities are then given by (our conventions here have the advantage of giving a proper ordering of string lengths, $`n_j>n_k`$, $`j>k`$) $`n_j=y_{i1}+(jm_i)y_i,v_j=(1)^{(n_j1)\zeta /\pi },m_ij<m_{i+1}.`$ (17) The total number of possible strings is $`N_s=m_l+1`$, and the index $`j`$ runs over the set $`1,\mathrm{},N_s`$. The real parameters $`\lambda _\alpha ^{n_j}`$ represent the centers of strings with length $`n_j`$ and parity $`v_j`$, and are hereafter noted as $`\lambda _\alpha ^j`$, $`\alpha =1,\mathrm{},M_j`$, where $`M_j`$ is the number of strings of length $`n_j`$ in the eigenstate under consideration. We therefore have the constraint $`_{j=1}^{N_s}n_jM_j=M`$. In a string configuration, many factors appearing in the Bethe equations become of the indeterminate form $`\delta /\delta `$. Remultiplying (3) for each member of a particular string gets rid of these factors, and allows one to rewrite the whole set of Bethe equations in terms of a reduced set involving only the string centers $`\lambda _\alpha ^j`$. Doing this, one finds (for $`N`$ even) the reduced set of Bethe-Takahashi equations TakahashiPTP46 $`\overline{e}_j^N(\lambda _\alpha ^j)=(1)^{M_j1}{\displaystyle \underset{(k,\beta )(j,\alpha )}{}}\overline{E}_{jk}(\lambda _\alpha ^j\lambda _\beta ^k),`$ (18) where $`\overline{e}_j(\lambda )=v_j{\displaystyle \frac{\mathrm{sinh}(\lambda +i\frac{\pi }{4}(1v_j)+in_j\zeta /2)}{\mathrm{sinh}(\lambda +i\frac{\pi }{4}(1v_j)in_j\zeta /2)}},\overline{E}_{jk}(\lambda )=\overline{e}_{|n_jn_k|}^{1\delta _{n_jn_k}}(\lambda )\overline{e}_{|n_jn_k|+2}^2(\lambda )\mathrm{}\overline{e}_{n_j+n_k2}^2(\lambda )\overline{e}_{n_j+n_k}(\lambda ).`$ (19) For the state classification and computation, it is preferable to work with the logarithmic form $`N\theta _j(\lambda _\alpha ^j){\displaystyle \underset{k=1}{\overset{N_s}{}}}{\displaystyle \underset{\beta =1}{\overset{M_k}{}}}\mathrm{\Theta }_{jk}(\lambda _\alpha ^j\lambda _\beta ^k)=2\pi I_\alpha ^j`$ (20) where $`I_\alpha ^j`$ is integer if $`M_j`$ is odd, and half-integer if $`M_j`$ is even. The dispersion kernels and scattering phases appearing here are $`\theta _j(\lambda )`$ $`=`$ $`2v_j\text{atan}\left[(\mathrm{tan}n_j\zeta /2)^{v_j}\mathrm{tanh}\lambda \right],`$ $`\mathrm{\Theta }_{jk}(\lambda )`$ $`=`$ $`(1\delta _{n_jn_k})\theta _{|n_jn_k|}(\lambda )+2\theta _{|n_jn_k|+2}(\lambda )+\mathrm{}+2\theta _{n_j+n_k2}(\lambda )+\theta _{n_j+n_k}(\lambda ).`$ (21) The energy and momentum of a string are given by $`E_\alpha ^j=J{\displaystyle \frac{\mathrm{sin}\zeta \mathrm{sin}n_j\zeta }{v_j\mathrm{cosh}2\lambda _\alpha ^j\mathrm{cos}n\zeta }},p_0(\lambda _\alpha ^j)=i\mathrm{ln}\left[{\displaystyle \frac{\mathrm{sinh}(\lambda _\alpha ^j+i\frac{\pi }{4}(1v_j)+in_j\zeta /2)}{\mathrm{sinh}(\lambda _\alpha ^j+i\frac{\pi }{4}(1v_j)in_j\zeta /2)}}\right]`$ (22) so the total momentum is again expressible in terms of the string quantum numbers as $`q=\pi T_{v=+}+{\displaystyle \frac{2\pi }{N}}{\displaystyle \underset{j=1}{\overset{N_s}{}}}{\displaystyle \underset{\alpha =1}{\overset{M_j}{}}}I_\alpha ^j\text{mod}2\pi `$ (23) in which $`T_{v=+}`$ is the total number of excitations with positive parity. It is known that the string hypothesis is not valid in absolute generality, meaning that not all eigenstates are described by a Bethe wavefunction with rapidities arranged in string patterns. For example, in the sector with two down spins of the isotropic antiferromagnet, it was demonstrated EsslerJPA25 that, for large lattice size $`N`$, there exist $`\text{O}(\sqrt{N})`$ pairs of extra real solutions replacing the corresponding expected two-strings (the total number of two-strings scaling like $`N`$). A fraction $`1/\sqrt{N}`$ of two-string states in the two down spin sector thus fail to be captured by the string hypothesis. For the anisotropic $`XXZ`$ model, again in the two down spin sector, it was argued IlakovacPRB60 that the number of violations of the string hypothesis remains finite for any $`\mathrm{\Delta }<1`$ as $`N`$ goes to infinity. Non-string states for the $`XXZ`$ model were also studied in \[FujitaJPA36, \]. In both cases, however, the total number of solutions to the Bethe equations given on the basis of the string hypothesis correctly reflects the dimensionality of the Hilbert space. If a string goes missing, it is replaced by a pair of real rapidities. The problem of counting the number of violations of the string hypothesis for larger numbers of overturned spins remains generally open. The one limit in which analytic progress can be made is at zero field, where $`M=N/2`$. There, it is possible to show WoynarovichJPA15 ; BabelonNPB220 that, in the thermodynamic limit, the lowest-lying excited states involving complex rapidities can only take the form of either complex conjugate pairs or self-conjugate quartets. Strings of length higher than two therefore seem to be excluded. We therefore expect to see a breakdown of higher-than-two strings as the magnetization decreases from its saturation value (i.e. the limit in which strings exist) down to zero. Since we do only a partial trace over the set of all intermediate states, including only a minuscule proportion of the whole set, we do not concern ourselves here with non-string states. This will have as a consequence that our results will be “at their worst” at zero field, where we only sum up states including two-string excitations, but will be essentially exact at higher fields. For each string state, we check explicitly that the deviations $`\delta `$ in (15) are exponentially small, and therefore only include these contributions if the string hypothesis is explicitly verified. In any case, we observe in general that strings of length higher than two give negligible contributions to the correlation functions. We therefore use the string hypothesis here as a general basis for our state classification scheme. The quality of our results will be confirmed later on with the use of sum rules. Our terminology for the state classification is as follows. First, we define a base as the set $`\{M_k\}`$, $`k=1,\mathrm{},N_s`$ of numbers of each string type present in an eigenstate. Many eigenstates share a same base, and these are obtained by choosing different quantum number configurations $`\{I_\alpha ^j\}`$, $`j=1,\mathrm{},N_s`$, $`\alpha =1,\mathrm{},M_j`$. The base itself defines the requirements on these quantum numbers. First, $`I_\alpha ^j`$ must be integer if $`M_j`$ is odd, and half-integer if $`M_j`$ is even. Second, they should be contained in the set $`\{I_{\mathrm{}}^j,I_{\mathrm{}}^j+1,\mathrm{},I_{\mathrm{}}^j\}`$, where $`I_{\mathrm{}}^j={\displaystyle \frac{1}{2\pi }}|N\theta _j(\mathrm{}){\displaystyle \underset{k=1}{\overset{N_s}{}}}(M_k\delta _{jk})\mathrm{\Theta }_{jk}(\mathrm{})|_j`$ (24) in which the notation $`a_j`$ means that we take the highest integer or half-integer less than or equal to the argument $`a`$, depending on the parity of $`M_j`$. The base choice completely specifies the right-hand side of this equation, and therefore the limits $`I_{\mathrm{}}^j`$. The total number of configurations sharing a given base is then given by $`D_{\{M\}}=_j\left(\begin{array}{c}2I_{\mathrm{}}^j+1\\ M_j\end{array}\right)`$. The total number of possible string states is then upon first examination given by summing this over all possible bases, $`D=_{\{M\}}D_{\{M\}}`$. This line of reasoning is used to argue for the completeness of the string states. However, we find that within a given base, there exist a number of inadmissible configurations. These correspond to configurations that are parity-symmetric, $`\{I_\alpha ^j\}=\{I_\alpha ^j\}`$ and that have a vanishing quantum number for at least one higher string, $`I_\alpha ^j=0`$ with even $`n_j`$, or that have two vanishing quantum numbers $`I_\alpha ^j,I_\beta ^k`$, $`jk`$, with $`n_j,n_k`$ odd. In the first case, there are rapidities exponentially close to $`i\zeta /2`$ (which represents an ill-defined limit when obtaining the Bethe equations), and in the second at least two exponentially close rapidities, meaning a wavefunction improperly defined by the straightforward Bethe Ansatz. The correct eigenstates should thus presumably be obtained by a proper limiting procedure. In view of the correctness of the string hypothesis state counting KirillovJPA30 , these must correspond to non-string states. We will readdress the question of state counting and non-string states in a separate publication. For our purposes here, however, completeness is an academic question, since we only need to do partial traces over intermediate states (albeit including strings). The presence of strings in the distribution of the rapidities $`\{\lambda \}`$ of an eigenstate renders the determinants used in (8), (11) degenerate, by which we mean that columns of the $`H,P,H^{}`$ and $`\mathrm{\Phi }`$ matrices become equal to leading order in the string deviations $`\delta `$. The matrix determinants must therefore be expanded to higher order, and column/row manipulations used to extract the first nonvanishing contribution. Overall, the form factors have proper limits. It is straightforward to show that the norm of the eigenstate with strings is given by the determinant of a reduced Gaudin matrix, $`|det\mathrm{\Phi }(\{\lambda \})|=\left[{\displaystyle \underset{j|n_j>1}{}}{\displaystyle \underset{\alpha }{}}{\displaystyle \underset{a=1}{\overset{n_j1}{}}}|(\delta _\alpha ^{n_j,a}\delta _\alpha ^{n_j,a+1})|\right]\left[|det\mathrm{\Phi }^{(r)}|+\text{O}(\delta )\right]`$ (25) of dimension equal to the number of strings in the state $`\{\lambda \}`$. Its matrix elements are defined as $`\mathrm{\Phi }_{(j,\alpha )(k,\beta )}^{(r)}=\delta _{jk}\delta _{\alpha \beta }\left(N{\displaystyle \frac{d}{d\lambda _\alpha ^j}}\theta _j(\lambda _\alpha ^j){\displaystyle \underset{(l,\gamma )(j,\alpha )}{}}{\displaystyle \frac{d}{d\lambda _\alpha ^j}}\mathrm{\Theta }_{jl}(\lambda _\alpha ^j\lambda _\gamma ^l)\right)+(1\delta _{jk}\delta _{\alpha \beta }){\displaystyle \frac{d}{d\lambda _\alpha ^j}}\mathrm{\Theta }_{jk}(\lambda _\alpha ^j\lambda _\beta ^k).`$ (26) In the case where the eigenstate described by the set $`\{\lambda \}`$ contains strings and the one described by $`\{\mu \}`$ does not, we can write explicit regular formulas for the elements of the matrices appearing in the longitudinal and transverse form factors. We find that the $`H`$ and $`P`$ matrix determinants for the longitudinal form factor can be reduced, in the presence of strings, to the determinants of reduced matrices up to corrections of order of the string deviations $`\delta `$, $`|det[H2P]|=|det[H^{(r)}2P^{(r)}]|+\text{O}(\delta ).`$ (27) The matrix components of $`H^{(r)}`$ and $`P^{(r)}`$, for matrix elements where $`\lambda _b`$ is part of an $`n`$-string, are given by $`H_{ab}^{(r)}`$ $`=`$ $`K_{ab}^{(n)},P_{ab}^{(r)}=0,b=1,\mathrm{},n1,`$ $`H_{ab}^{(r)}`$ $`=`$ $`N^{(n)}{\displaystyle \underset{j=0}{\overset{n}{}}}{\displaystyle \frac{G_j^{(n)}G_{j+1}^{(n)}}{F_j^{(n)}F_{j+1}^{(n)}}}\left([\delta _{j,0}+\delta _{j,n}1]L_{aj}^{(n)}+[\delta _{j,0}+\delta _{j,n}]K_{aj}^{(n)}\right),`$ $`P_{ab}^{(r)}`$ $`=`$ $`N^{(n)}{\displaystyle \frac{1}{\mathrm{sinh}^2\mu _a+\mathrm{sin}^2\zeta /2}}{\displaystyle \underset{j=1}{\overset{n}{}}}{\displaystyle \frac{G_j^{(n)}}{F_j^{(n)}}},b=n,`$ (28) where $`F_b^{(n)}`$ $`=`$ $`{\displaystyle \underset{kb}{}}\mathrm{sinh}\left[\lambda _k\lambda _b\right],G_b^{(n)}={\displaystyle \underset{j=1}{\overset{M}{}}}\mathrm{sinh}\left[\mu _j\lambda _b\right],`$ $`K_{ab}^{(n)}`$ $`=`$ $`\left[\mathrm{sinh}(\mu _a\lambda _b)\mathrm{sinh}(\mu _a\lambda _b+i\zeta )\right]^1,L_{ab}^{(n)}={\displaystyle \frac{d}{d\lambda }}K_{ab}^{(n)},`$ $`N^{(n)}`$ $`=`$ $`F_0^{(n)}F_1^{(n)}{\displaystyle \frac{_{j=2}^nG_j^{(n)}}{G_n^{(n)}}}.`$ (29) in which we have extended the definitions of $`\lambda _b`$ in (15) to include $`a=0`$ and $`a=n+1`$ elements for notational convenience. In the last formula above, the product is defined as unity for $`n=1`$. Note that these remain $`M`$-dimensional matrices (the dimension doesn’t change here, in contrast to the Gaudin case). A similar procedure can be employed to treat the other form factor: the first $`M1`$ columns of the reduced transverse form factor matrix $`H^{(r)}`$ are given by the same expressions as for $`H^{(r)}`$(bearing in mind that the set $`\{\lambda \}`$ then has $`M1`$ elements), and the $`M`$-th column remains unchanged from equation (12) in the presence of strings in the set $`\{\lambda \}`$. Therefore, if $`\{\lambda \}`$ contains strings and $`\{\mu \}`$ does not, we find that the string deviations $`\delta `$ cancel out in the product for the longitudinal form factor, and that (with exponentially small corrections of order $`\delta `$) we can write $`|\{\mu \}|S_q^z|\{\lambda \}|^2`$ $`=`$ $`{\displaystyle \frac{N}{4}}\delta _{q,q_{\{\lambda \}}q_{\{\mu \}}}{\displaystyle \underset{j=1}{\overset{M}{}}}\left|{\displaystyle \frac{\mathrm{sinh}(\mu _ji\zeta /2)}{\mathrm{sinh}(\lambda _ji\zeta /2)}}|^2{\displaystyle \underset{jk=1}{\overset{M}{}}}\right|\mathrm{sinh}(\mu _j\mu _k+i\zeta )|^1\times `$ (30) $`\times {\displaystyle \underset{jk=1,\lambda _j\lambda _ki\zeta }{\overset{M}{}}}|\mathrm{sinh}(\lambda _j\lambda _k+i\zeta )|^1{\displaystyle \frac{\left|det[𝐇^{(r)}(\{\mu \},\{\lambda \})2𝐏^{(r)}(\{\mu \},\{\lambda \})]\right|^2}{\left|det𝚽(\{\mu \})det𝚽^{(r)}(\{\lambda \})\right|}}`$ The transverse form factor similarly reduces to $`|\{\mu \}|S_q^{}|\{\lambda \}|^2`$ $`=`$ $`N\delta _{q,q_{\{\lambda \}}q_{\{\mu \}}}|\mathrm{sin}\zeta |{\displaystyle \frac{_{j=1}^M\left|\mathrm{sinh}(\mu _ji\zeta /2)\right|^2}{_{j=1}^{M1}\left|\mathrm{sinh}(\lambda _ji\zeta /2)\right|^2}}{\displaystyle \underset{jk=1}{\overset{M}{}}}|\mathrm{sinh}(\mu _j\mu _k+i\zeta )|^1\times `$ (31) $`\times {\displaystyle \underset{jk=1,\lambda _j\lambda _ki\zeta }{\overset{M1}{}}}|\mathrm{sinh}(\lambda _j\lambda _k+i\zeta )|^1{\displaystyle \frac{\left|det𝐇^{(r)}(\{\mu \},\{\lambda \})\right|^2}{\left|det𝚽(\{\mu \})det𝚽^{(r)}(\{\lambda \})\right|}}`$ For proper string states, all the above functions and determinants are nondegenerate. In the next section, we use these results to numerically compute the form factors, and sum them to obtain the dynamical correlation functions. ## IV Dynamical spin-spin correlations The strategy to follow is now clear. We compute the $`S^{zz}`$ and $`S^+`$ structure factors by directly summing the terms on the right-hand side of equation (7) over a judiciously chosen subset of eigenstates. The momentum delta functions are broadened to width $`ϵ1/N`$ using $`\delta _ϵ(x)=\frac{1}{\sqrt{\pi }ϵ}e^{x^2/ϵ^2}`$ in order to obtain smooth curves. We scan through the eigenstates in the following order. First, we observe that the form factors of the spin operators between the ground state and an eigenstate $`\{\lambda \}`$ are extremely rapidly decreasing functions of the number of holes that need to be inserted in the configuration of the lowest-energy state (in the same base) in order to obtain the configuration $`\{I\}`$ corresponding to $`\{\lambda \}`$. We therefore scan through all bases and configurations for increasing number of holes, starting from one-hole states for $`S^{zz}`$, and zero-hole states for $`S^+`$. Although the number of possible configurations for fixed base and number of holes is a rapidly increasing function of the number of holes, we find that the total contributions for fixed bases also rapidly decrease for increasing hole numbers. We therefore limit ourselves to states with up to three holes, corresponding to up to six-particle excitations. We can quantify the quality of the present computational method by evaluating the sum rules for the longitudinal and transverse form factors. Namely, by integrating over momentum and frequency, we should saturate the values $`{\displaystyle _{\mathrm{}}^{\mathrm{}}}{\displaystyle \frac{d\omega }{2\pi }}{\displaystyle \frac{1}{N}}{\displaystyle \underset{q}{}}S^{zz}(q,\omega )={\displaystyle \frac{1}{4}}S^z^2={\displaystyle \frac{1}{4}}\left[1(1{\displaystyle \frac{2M}{N}})^2\right]`$ (32) $`{\displaystyle _{\mathrm{}}^{\mathrm{}}}{\displaystyle \frac{d\omega }{2\pi }}{\displaystyle \frac{1}{N}}{\displaystyle \underset{q}{}}S^+(q,\omega )={\displaystyle \frac{1}{2}}S^z={\displaystyle \frac{M}{N}}.`$ (33) In Fig. 1, we plot the longitudinal structure factor as a function of momentum and frequency for anisotropy $`\mathrm{\Delta }=0.25`$, for four values of the magnetization. Fig. 3 contains the transverse structure factor for the same anisotropy and magnetizations, whereas Figs. 2 and 4 give plots for another value of anisotropy, $`\mathrm{\Delta }=0.75`$. For all intermediate states involving strings, we explicitly check that the deviations from the string hypothesis are small. We find in general that states involving strings of length higher that two are admissible solutions to the Bethe equations for high enough magnetizations. At zero field, only two-string states have exponentially small deviations $`\delta `$, and all higher-string states must be discarded, consistent with \[WoynarovichJPA15, ; BabelonNPB220, \]. The relative contributions to the structure factors from different bases is very much dependent on the system size, the anisotropy, and the magnetization. In general, we find that two- and four-particle contributions are sufficient to saturate well over 90% of the sum rules in all cases, for system sizes up to $`N=200`$. Interestingly, however, we find that string states also contribute noticeably in many cases. For example, in Fig. 5, we plot the zero-field transverse structure factor contributions coming from intermediate states with one string of length two and up to three holes. Around six or seven percent of the weight is accounted for by these states, and similar or somewhat lower figures are found in other cases. Strings of length higher than two do not contribute significantly. For example, we find only around 5.7e-8 % of the sum rule from states with one string of length three, for the longitudinal structure factor for $`\mathrm{\Delta }=0.25`$ at $`M=N/4`$ with $`N=128`$. For $`\mathrm{\Delta }=0.75`$, we find 6.3e-7 %. For the transverse correlators, the figures are 2.3e-12 % and 3.1e-12 %. Even though these numbers would increase if we could go to larger system sizes, we do not expect them to ever become numerically significant. The imperfect saturation of the sum rules that we obtain in general can be ascribed either to higher states in the hierarchy which are not included in our partial summations, or states that are in principle included, but which are rejected in view of their deviations from the string hypothesis. As the proportion of excluded string states to allowable ones can be rather large (ranging anywhere from zero to fifty percent), we believe the latter explanation to be the correct one. We will attempt to include these non-string states in a future improvement of our method. ## V Real space- and time-dependent correlators From our results covering the whole Brillouin zone and frequency space, it is straightforward to obtain space-time dependent correlation functions by inverse Fourier transform: $`S_{j+1}^a(t)S_1^{\overline{a}}(0)_c={\displaystyle \frac{1}{N}}{\displaystyle \underset{\alpha }{}}|GS|S_{q_\alpha }^a|\alpha |^2e^{iq_\alpha ji\omega _\alpha t}.`$ (34) In the present section, we compare these results to known exact results for equal-time correlators at short distance, and to the large-distance asymptotic form obtained from conformal field theory. This comparison can only be made at zero field, where both sets of results are known exactly. ### V.1 Small distances and equal time It is known that correlation functions of spin operators on the lattice can be represented in general as multidimensional integrals JimboPLA168 ; JimboBOOK ; JimboJPA29 ; KitanineNPB567 ; KitanineNPB641 ; KitanineJPA35 ; Boosh0412191 . For example, the spin-spin correlation function at distance $`j`$ is exactly represented by a $`j`$-fold integral. These are unfortunately rather difficult to evaluate either analytically or numerically, and exact answers are only known for small $`j`$ or in the asymptotic limit $`j\mathrm{}`$. Recently, large simplifications were obtained for integral representations of the emptiness formation probability BoosJPA34 ; BoosJPA35 ; BoosNPB658 , and for third-neighbour correlations SakaiPRE67 in the case of the isotropic Heisenberg chain in zero field. Results for the anisotropic chain in zero field were subsequently obtained in \[KatoJPA36, ; TakahashiJPSJ73, ; KatoJPA37, \], yielding exact results for equal-time correlators of spin operators on up to four adjacent sites. Explicitly, these correlators read KatoJPA37 $`S_j^zS_{j+1}^z`$ $`=`$ $`F_{++}^{++}{\displaystyle \frac{1}{4}},S_j^{}S_{j+1}^+=F_+^+,S_j^zS_{j+2}^z=2F_{+++}^{+++}2F_{++}^{++}+{\displaystyle \frac{1}{4}},S_j^{}S_{j+2}^+=2F_{++}^{++},`$ $`S_j^zS_{j+3}^z`$ $`=`$ $`2F_{++++}^{++++}2F_{++}^{++}3F_{++}^{++}+{\displaystyle \frac{3}{4}},S_j^{}S_{j+2}^+=2F_{+++}^{+++}+2F_{++}^{++},`$ (35) where all blocks $`F`$, representing correlation functions of $`2\times 2`$ fundamental matrices composed of vanishing elements except for one, are given explicitly as polynomials of the two one-dimensional integrals $`\zeta _n(j)={\displaystyle _{\mathrm{}i\pi /2}^{\mathrm{}i\pi /2}}𝑑x{\displaystyle \frac{1}{\mathrm{sinh}x}}{\displaystyle \frac{\mathrm{cosh}\eta x}{\mathrm{sinh}^j\eta x}},\zeta _n^{}(j)={\displaystyle _{\mathrm{}i\pi /2}^{\mathrm{}i\pi /2}}𝑑x{\displaystyle \frac{1}{\mathrm{sinh}x}}{\displaystyle \frac{}{\eta }}{\displaystyle \frac{\mathrm{cosh}\eta x}{\mathrm{sinh}^j\eta x}}.`$ (36) The explicit form of these polynomials can be found for the $`XXZ`$ case in the second appendix of paper \[KatoJPA37, \]. We compare these results to our own values, obtained from Fourier transform of the form factors, in tables 1 and 2. The deviations between the results are overall even smaller than can be expected on the basis of our slight deviations from perfect sum rule saturation. For the particular case of $`\mathrm{\Delta }=0.5`$, exact results at zero field were also recently obtained up to distances of eight sites Kitanineh0506114 . We compare these results to our form factor calculations in table 3. ### V.2 The scaling limit The continuum limit of the anisotropic $`XXZ`$ chain in the gapless regime $`1<\mathrm{\Delta }1`$ is described by a Gaussian conformal field theory LutherPRB12 ; KadanoffAP121 . Expressions for the lattice Hamiltonian and for local spin operators as asymptotic expansions of local scaling fields were obtained to subleading order in \[LukyanovPRB59, ; LukyanovNPB654, \], whose results we use here as comparison. For large distance, the correlators explicitly read $`S_{j+l}^{}(t)S_j^+(0)_c`$ $``$ $`(1)^l{\displaystyle \frac{A/2}{(l_+l_{})^{\frac{\eta }{2}}}}\left\{1{\displaystyle \frac{B}{(l_+l_{})^{\frac{2}{\eta }2}}}+\mathrm{}\right\}{\displaystyle \frac{\stackrel{~}{A}/2}{(l_+l_{})^{\frac{\eta }{2}+\frac{1}{2\eta }}}}\left\{{\displaystyle \frac{1}{2}}\left({\displaystyle \frac{l_+}{l_{}}}+{\displaystyle \frac{l_{}}{l_+}}\right)+{\displaystyle \frac{\stackrel{~}{B}}{(l_+l_{})^{\frac{1}{\eta }1}}}+\mathrm{}\right\},`$ $`S_{j+l}^z(t)S_j^z(0)_c`$ $``$ $`{\displaystyle \frac{1}{4\pi ^2\eta l_+l_{}}}\left\{{\displaystyle \frac{1}{2}}\left({\displaystyle \frac{l_+}{l_{}}}+{\displaystyle \frac{l_{}}{l_+}}\right)+{\displaystyle \frac{\stackrel{~}{B}_z}{(l_+l_{})^{\frac{2}{\eta }2}}}\left(1+{\displaystyle \frac{2\eta }{4(1\eta )}}\left({\displaystyle \frac{l_+}{l_{}}}+{\displaystyle \frac{l_{}}{l_+}}\right)+\mathrm{}\right)\right\}+`$ (37) $`+{\displaystyle \frac{(1)^lA_z/4}{(l_+l_{})^{\frac{1}{2\eta }}}}\left\{1{\displaystyle \frac{B_z}{(l_+l_{})^{\frac{1}{\eta }1}}}+\mathrm{}\right\}`$ with $`l_\pm =l\pm t/ϵ`$. To compare with our results, we use a conformal transformation to finite size, so that the chiral distances become $`l_\pm =\frac{N}{\pi }\mathrm{sin}(\frac{\pi }{N}(l\pm t))`$. Here, the anisotropy appears in the parameter $`\eta =1\frac{\zeta }{\pi }`$, and the amplitudes $`A,\stackrel{~}{A},A_z,B,\stackrel{~}{B},B_z`$ are known exactly for zero field, e.g. $`A={\displaystyle \frac{1}{2(1\eta )^2}}\left[{\displaystyle \frac{\mathrm{\Gamma }(\frac{\eta }{22\eta })}{2\sqrt{\pi }\mathrm{\Gamma }(\frac{1}{22\eta })}}\right]^\eta \mathrm{exp}\left\{{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dt}{t}}\left({\displaystyle \frac{\mathrm{sinh}(\eta t)}{\mathrm{sinh}t\mathrm{cosh}((1\eta )t)}}\eta e^{2t}\right)\right\}`$ (38) (the explicit expressions for the other amplitudes, which we do not reproduce here, can be found in \[LukyanovNPB654, \]). Tables 4 and 5 contain comparisons of the numerical values of the equal-time spin-spin correlation functions computed using the CFT aymptotic expansions and our form factor calculations. Fig. 6 contains plots of the results as a function of lattice distance. Once again, the deviations are even smaller than should be expected, considering the imperfect sum rule saturation the the fact that these remain finite size comparisons. The only substantial deviations occur at small distances, where the conformal approach yields incorrect results (as a comparison to the data from the previous subsection shows). ## VI Conclusion The existence of determinant representations for lattice form factors of local spin operators makes it possible to compute dynamical correlation functions to a high degree of accuracy, and we have here attempted to push the results of \[Caux0502365, \] further by including the remaining spin-spin correlators, studying their general field dependence, including contributions from intermediate states containing complex solutions to the Bethe equations in the presence of strings, and computing explicitly real space- and time-dependent correlators. Although intermediate states with strings of length two do indeed give non-negligible contributions to the sum rules, we have found that strings of any higher length have form factors which are much too small to be of significance. We have demonstrated the reliability of our results by performing comparisons with exact integral representations for equal time, small-distance correlators at zero field, and exact formulae for the large-distance asymptotics (again at zero field). We do wish to emphasize that the present method, unlike these calculations, is not limited to zero field. Also, unlike other numerical methods based on the Density Matrix Renormalization Group (DMRG), it is by construction directly applicable to dynamics. In other words, we are able to numerically compute all spin-spin correlators either as a function of momentum and frequency, or space and time separation, for any value of magnetization, for system sizes well beyond the reach of any other numerical method. ###### Acknowledgements. J.-S. C. acknowledges support from CNRS and the Stichting voor Fundamenteel Onderzoek der Materie (FOM). J.-M. M. is supported by CNRS, and acknowledges support from the EUCLID European network.
warning/0506/quant-ph0506068.html
ar5iv
text
# Conditional probabilities and density operators in quantum modeling ## I Introduction A recent proof confirms what to some will seem a commonplace: the equations of quantum mechanics are separated by a logical gap from their application to describing experiments with devices, so that choosing equations to describe devices involves an irreducible element of judgment 0404113 . This finding impacts a core question of interpretation: what does quantum mechanics describe leggett ? Although a wide variety of interpretations accept the assumptions on which the proof depends, the proof tells us that quantum mechanics by itself describes nothing, but instead offers a mathematically articulated language that people speak to describe what they see or expect or think possible in experiments ams ; JOptB . With the recognition of a logical gap between the equations and experiments comes an opportunity to examine relations among some purely mathematical entities—probabilities, density operators, partial traces—separated out from the choices and judgments necessary to apply them to describing experiments. Based on this examination, I will show uses of conditional density operators defined in relation to the trace rule by which quantum mechanics generates probabilities from density operators and positive operator-valued measures. The next section reviews the expression of joint and conditional probabilities by tensor products of projection-valued measures. Section III generalizes from projection-valued measures to positive operator-valued measures (POVMs) with a proposition that lifts density operators to an extended Hilbert space associated with Neumark’s theorem. Diagrams show how tensor products of POVMs (but not generic operator products) express joint probabilities, leading to conditional density operators useful in modeling measurements of composite systems. Section IV shows how choices in the application of quantum mathematics to the description of devices allows a single, discrete measure space to be applied in describing diverse situations. Section V offers two applications of the conditional probabilities and partial traces: (1) a demonstration that the use of reduced states in quantum mechanics requires no postulate about “effects of a measurement on a state”; and (2) a demonstration of the equivalence of two forms of quantum key distribution (QKD) qkd , which I term “transmitted-state BB84” and “entangled-state BB84” BB84 with the result that a known formulation slutsky for studying individual eavesdropping attacks against transmitted-state BB84 applies also to entangled-state BB84. ## II Probabilities, Operator-valued measures, and quantum modeling By a probability measure I mean a completely additive set function kol , or, in more modern words, a positive measure mackey ; rudin of total measure 1. Let $`\mathrm{\Omega }`$ be a topological space; let $`(\mathrm{\Omega })`$ be the $`\sigma `$-algebra of measurable subsets in $`\mathrm{\Omega }`$, making $`\mathrm{\Omega }`$ into a measurable space. Let $`\mu `$ denote any probability measure on a fixed $`(\mathrm{\Omega })`$. For any measurable sets $`X`$ and $`Y`$, with $`\mu (Y)>0`$, the conditional probability of $`X`$ given $`Y`$ is defined kol as: $$\mu (X|Y)\stackrel{\mathrm{def}}{=}\mu (XY)/\mu (Y).$$ (1) Let $`()`$ denote the set of bounded, linear operators on a Hilbert space $``$. With $`(\mathrm{\Omega })`$ as above, a positive operator-valued measure (POVM) is any function $$M:(\mathrm{\Omega })()$$ (2) satisfying: (1) $`M(\mathrm{})=0`$, $`M(\mathrm{\Omega })=\mathrm{𝟏}_{}`$; (2) each $`M(X)`$ is self-adjoint and non-negative; (3) if $`XY=0`$ then $`M(XY)=M(X)+M(Y)`$; and (4) for every $`|u,|v`$, the set function $`M_{u,v}`$ defined by $`M_{u,v}(X)=v|M(X)|u`$ is a complex measure on $`(\mathrm{\Omega })`$. A projection-valued measure, sometimes called a projective resolution of the identity rudin , is a special case of a POVM, denoted here by $`E`$ in place of $`M`$, that satisfies the above requirements, and, in addition: Each $`E(X)`$ is a self-adjoint projection $`(`$so $`E^2(X)=E(X)`$$`)`$; (3) $`E(XY)=E(X)E(Y).`$ (4) Note that this last property implies $$[E(X),E(Y]\stackrel{\mathrm{def}}{=}E(X)E(Y)E(Y)E(X)=0.$$ (5) I take a density operator on $``$ to be any positive, self-adjoint trace-class operator $`\rho ()`$ such that $`\text{Tr}(\rho )=1`$, where Tr denotes the trace vN . If $`\rho `$ is a density operator on $``$, so is $`U\rho U^{}`$ where $`U`$ is any unitary operator on $``$. For any POVM defined for $`(\mathrm{\Omega })`$ and $``$, along with any density operator $`\rho `$ on $``$ and any unitary operator $`U`$ on $``$, a probability measure on $`(\mathrm{\Omega })`$ is defined by $$\mu (\rho ,U,M;X)=\text{Tr}[U\rho U^{}M(X)].$$ (6) In the modeling of experiments one is interested in a family of such measures, corresponding to various choices of $`\rho `$, $`U`$, and $`M`$. Often one skips the listing of these “parameters” and writes $`\mathrm{Pr}(X)`$ in place of $`\mu (\rho ,U,M;X)`$. Feller speaks of $`\mathrm{\Omega }`$ as a “sample space” and of $`(\mathrm{\Omega })`$ as “the set of events” feller , inviting us to think of a “probability that a sample point falls in an event set X”; and, correspondingly, to think of $`\mu (X|Y)`$ as a conditional probability that a sample point falls in $`X`$ given that it falls in $`Y`$. (Without drawing a distinction between a sample point and an event, Dirac dirac speaks of a “result” while Peres peres speaks of an “outcome.”) In quantum physics we speak of a preparation of a state $`\rho `$, a time evolution $`U`$, and a measurement expressed by $`M`$. Quantum decision theory helstrom adjoins to this story of “quantum probability” a “classical probability” in the choice of the state $`\rho `$ prepared, making two kinds of events. To distinguish these kinds, I rename the event above a measurement event and join it to a state event expressing preparation of the state. States are usually thought of as selected from a discrete space of possible states, in which case the state event can be denoted simply by the state chosen, resulting in a compound event consisting of a measurement event $`X`$ (as before) together with a state event $`\rho `$. The marginal probability for choosing $`\rho `$ can be denoted $`\mathrm{Pr}(\rho )`$. With this convention, decision problems are formulated in terms of a (joint) measure $$\mu (U,M;\rho ,X)=\mathrm{Pr}(\rho )\text{Tr}[U\rho U^{}M(X)],$$ (7) where on the left-hand side of the equation $`\rho `$ has moved from the ‘parameter side’ of the semicolon to the ‘variable side.’ If $`U`$ and $`M`$ are understood, one can write $`\mathrm{Pr}(\rho ,X)`$ in place of $`\mu (U,M;\rho ,X)`$. (Models that go further by randomizing the choices of $`M`$ and/or $`U`$ can be found, but not in this report.) By quantum modeling I mean stating probabilities in the form of Eqs. (6) or (7) (together with probabilities derived from these by Bayes’ rule) as mathematical language by which to ask questions and make statements pertaining to a set of trials of devices in a laboratory experiment 0404113 ; ams . Examples of devices are lasers, lenses, and detectors. In this language a trial is necessarily described as consisting of “preparing a state” and “measuring a state,” in some cases interspersing these with a temporal evolution $`U`$. In quantum modeling, one speaks of various conditional probabilities, conditioned on whatever types of events are expressed in a model; hence there can be conditioning not only on measurement events but also on state events: 1. Understanding some $`\rho `$, $`M`$, and $`U`$, one can speak of the conditional probability $`\mathrm{Pr}(X|Y)`$ of a sample point being in measurement event set $`X`$, given it is in set $`Y`$. 2. Understanding some $`M`$ and $`U`$, one can speak of the conditional probability of a measurement event $`X`$, given the state is $`\rho `$. For example, from Eq. (7) and the definition of a conditional probability, we see what in Eq. (6) appeared as a probability becomes in the context of decision theory a species of conditional probability: $$\mathrm{Pr}(X|\rho )=\frac{\mathrm{Pr}(\rho ,X)}{\mathrm{Pr}(\rho )}=\text{Tr}[U\rho U^{}M(X)].$$ (8) Variations on both of these appear below. ### II.1 Conditional density operators for a single projection-valued measure For the rest of this section and all of Sec. III, I subsume $`U\rho U^{}`$ into $`\rho `$. By virtue of Eqs. (1), (4) and (5), conditional probabilities of a sample point being in one measurement event given that the point is in another measurement event fit in neatly with any single projection-valued measure $`E`$: $`(X,Y(\mathrm{\Omega })\text{ with }\mathrm{Pr}(Y)>0)\mathrm{Pr}(X|Y)`$ $`=`$ $`{\displaystyle \frac{\text{Tr}[\rho E(X)E(Y)]}{\text{Tr}[\rho E(Y)]}}={\displaystyle \frac{\text{Tr}[E(Y)\rho E(Y)E(X)]}{\text{Tr}[\rho E(Y)]}},`$ (9) which allows us to define a “conditional density operator” $$\rho |_Y\stackrel{\mathrm{def}}{=}\frac{E(Y)\rho E(Y)}{\text{Tr}[\rho E(Y)]},$$ (10) with the property that $$(X)\mathrm{Pr}(X|Y)=\text{Tr}[\rho |_YE(X)].$$ (11) Similarly one constructs $`\rho |_X`$ such that $$(X)\mathrm{Pr}(Y|X)=\text{Tr}[\rho |_XE(Y)].$$ (12) It is easy to check that the conditional density operators $`\rho |_Y`$ and $`\rho |_X`$ are non-negative, self-adjoint, and have unit trace. The expression of a conditional probability by a conditional density operator depends on the expression of a joint probability by a product of projections, as illustrated by the following commutative diagram (where commutative means that alternative ways to compose mappings arrive at the same thing): X XY Y E(X) E(X)E(Y) E(Y).𝑋fragments fragmentsXYfragments 𝑌 missing-subexpression missing-subexpression fragmentsE(X)fragments fragments E(X)E(Y)fragments fragmentsE(Y){\begin{tabular}[]{ccccc}$X$&\ $\longmapsto$&$X\cap Y$&\ \raise 5.5pt\hbox{\rotatebox{180.0}{$\longmapsto$}}&$Y$\\[-12.0pt] \rotatebox{270.0}{$\longmapsto$}&&\rotatebox{270.0}{$\longmapsto$}&&\rotatebox{270.0}{$\longmapsto$}\\ $E(X)$&\ $\longmapsto$&\ $E(X)E(Y)$&\ \raise 5.5pt\hbox{\rotatebox{180.0}{$\longmapsto$}}&$E(Y)$\end{tabular}.} (13) Diagram 1: $`E(XY)=E(X)E(Y)`$. Although most of the conditional density operators to be introduced below are constructed using partial traces, $`\rho |_Y`$ as defined by Eq. (10) involves no partial trace and so is not a reduced density operator, which shows that the concept of a conditional density operator is distinct from the concept of a reduced density operator. Note: The conditional probabilities above are just probabilities that a measurement sample point that falls in a set $`Y`$ also falls in a set $`X`$; this has nothing to do with “consecutive measurements,” nor with so-called “state reductions” or “collapse of a wave function” postulated by Dirac dirac and von Neumann vN . ### II.2 Tensor products of projection-valued measures Consider a special case that arises in the modeling of systems viewed as composites of subsystems $`A`$ and $`B`$, so that $`\mathrm{\Omega }`$ $`=`$ $`\mathrm{\Omega }_A\times \mathrm{\Omega }_B\text{(cartesian product)},`$ $``$ $``$ $`_{AB}=_A_B\text{(tensor product)},`$ $`E`$ $`=`$ $`E_AE_B\text{ with }E_A\text{ on }_A\text{ and }E_B\text{ on }_B,`$ $`\rho `$ $`=`$ $`\rho _{AB}(_{AB}),`$ $`U`$ $`=`$ $`\mathrm{𝟏}_{AB}.`$ (14) The tensor product of projection-valued measures is defined by: $$(X_A(\mathrm{\Omega }_A),Y_B(\mathrm{\Omega }_B))(E_AE_B)(X_A\times X_B)=E_A(X_A)E_B(Y_B).$$ (15) Then we have $`(X_A\times \mathrm{\Omega }_B)(\mathrm{\Omega }_A\times Y_B)`$ $`=`$ $`(X_A\times Y_B),`$ (16) $`(E_AE_B)[(X_A\times \mathrm{\Omega }_B)(\mathrm{\Omega }_A\times Y_B)]`$ $`=`$ $`(E_AE_B)(X_A\times Y_B)=E_A(X_A)E_B(Y_B),`$ (17) and the following diagram commutes: XA XA×ΩB (XA×ΩB)(ΩA×YB) ΩA×YB YB EA(XA) EA(XA)𝟏B EA(XA)EB(YB) 𝟏AEB(YB) EB(YB).fragmentsX𝐴fragments fragmentsX𝐴Ω𝐵fragments fragments(X𝐴Ω𝐵)(Ω𝐴Y𝐵) fragmentsΩ𝐴Y𝐵 fragmentsY𝐵 missing-subexpression missing-subexpression missing-subexpression missing-subexpression fragmentsE𝐴(X𝐴)fragments fragmentsE𝐴(X𝐴)tensor-product1𝐵fragments fragmentsE𝐴(X𝐴)tensor-productE𝐵(Y𝐵) fragments1𝐴tensor-productE𝐵(Y𝐵) fragmentsE𝐵(Y𝐵){\begin{tabular}[]{ccccccccc}$X_{A}\mskip-4.0mu$&\ $\longmapsto$&$\mskip-4.0muX_{A}\times\Omega_{B}\mskip-4.0mu$&\ $\longmapsto$&$(X_{A}\times\Omega_{B})\cap(\Omega_{A}\times Y_{B})\mskip-4.0mu$&\raise 5.5pt\hbox{\rotatebox{180.0}{$\longmapsto$}}&$\mskip-4.0mu\Omega_{A}\times Y_{B}\mskip-4.0mu$&\raise 5.5pt\hbox{\rotatebox{180.0}{$\longmapsto$}}&$\mskip-4.0muY_{B}$\\[-12.0pt] \rotatebox{270.0}{$\longmapsto$}&&\rotatebox{270.0}{$\longmapsto$}&&\rotatebox{270.0}{$\longmapsto$}&&\rotatebox{270.0}{$\longmapsto$}&&\rotatebox{270.0}{$\longmapsto$}\\ $E_{A}(X_{A})\mskip-4.0mu$&\ $\longmapsto$&$E_{A}(X_{A})\otimes\bm{1}_{B}\mskip-4.0mu$&\ $\longmapsto$&$\mskip-4.0muE_{A}(X_{A})\otimes E_{B}(Y_{B})\mskip-4.0mu$&\raise 5.5pt\hbox{\rotatebox{180.0}{$\longmapsto$}}&$\bm{1}_{A}\otimes E_{B}(Y_{B})\mskip-4.0mu$&\raise 5.5pt\hbox{\rotatebox{180.0}{$\longmapsto$}}&$E_{B}(Y_{B})$\end{tabular}.} (18) Diagram 2: Tensor product of projection-valued measures. For this specialization of Diagram 1, it follows from Eqs. (9) and (17) that $`\mathrm{Pr}(X_A|Y_B)`$ $``$ $`\mathrm{Pr}(X_A\times \mathrm{\Omega }_B|\mathrm{\Omega }_A\times Y_B)`$ (19) $`=`$ $`{\displaystyle \frac{\text{Tr}_{AB}\{[\mathrm{𝟏}_AE_B(Y_B)]\rho _{AB}[\mathrm{𝟏}_AE_B(Y_B)](E_A(X_A)\mathrm{𝟏}_B)\}}{\text{Tr}_{AB}\{\rho _{AB}[\mathrm{𝟏}_AE_B(Y_B)]\}}},`$ where the subscript $`AB`$ on the trace indicates the trace over $`_{AB}`$. From this follow two distinct ways to define a conditional density operator: 1. Eq. (12) implies: $`\mathrm{Pr}(X_A|Y_B)`$ $`=`$ $`\text{Tr}_{AB}[\rho _{AB}^{(1)}|_{Y_B}(E_A(X_A)\mathrm{𝟏}_B)],`$ (20) with $`\rho _{AB}^{(1)}|_{Y_B}`$ $`\stackrel{\mathrm{def}}{=}`$ $`{\displaystyle \frac{[\mathrm{𝟏}_AE_B(Y_B)]\rho _{AB}[\mathrm{𝟏}_AE_B(Y_B)]}{\text{Tr}_{AB}\{\rho _{AB}[\mathrm{𝟏}_AE_B(Y_B)]\}}}.`$ (21) 2. The alternative definition takes advantage of partial traces to obtain a conditional density operator as a reduced density operator: $$\mathrm{Pr}(X_A|Y_B)=\frac{\text{Tr}_A\text{Tr}_B\{[\mathrm{𝟏}_AE_B(Y_B)]\rho _{AB}[\mathrm{𝟏}_AE_B(Y_B)](E_A(X_A)\mathrm{𝟏}_B)\}}{\text{Tr}_{AB}\{\rho _{AB}[\mathrm{𝟏}_AE_B(Y_B)]\}},$$ (22) which, with $`XX_A`$ and $`YY_B`$ understood, can be written with less clutter as $`\mathrm{Pr}(X|Y)`$ $`=`$ $`{\displaystyle \frac{\text{Tr}_A\text{Tr}_B\{[\mathrm{𝟏}_AE_B(Y)]\rho _{AB}[\mathrm{𝟏}_AE_B(Y)](E_A(X)\mathrm{𝟏}_B)\}}{\text{Tr}_{AB}\{\rho _{AB}[\mathrm{𝟏}_AE_B(Y)]\}}}`$ (23) $`=`$ $`\text{Tr}_A[\rho _A^{(2)}|_Y(E_A(X)\mathrm{𝟏}_B)],`$ with $`\rho _A^{(2)}|_Y`$ $`\stackrel{\mathrm{def}}{=}`$ $`{\displaystyle \frac{\text{Tr}_B\{[\mathrm{𝟏}_AE_B(Y)]\rho _{AB}[\mathrm{𝟏}_AE_B(Y)]\}}{\text{Tr}_{AB}\{\rho _{AB}[\mathrm{𝟏}_AE_B(Y)]\}}}`$ (24) $`=`$ $`{\displaystyle \frac{\text{Tr}_B\{\rho _{AB}[\mathrm{𝟏}_AE_B(Y)]\}}{\text{Tr}_{AB}\{\rho _{AB}[\mathrm{𝟏}_AE_B(Y)]\}}}.`$ I stretch notation to allow $`\text{Tr}_A`$ to indicate either the partial trace over the factor $`_A`$ of the tensor-product Hilbert space $`_A_B`$ or the trace over just the Hilbert space $`_A`$. Properties of the partial trace used in these equations are reviewed in Appendix A; for example, Eq. (68) assures us that the last expression in Eq. (24) is a self-adjoint operator on $`_A`$. When this mathematics of tensor products is applied to model a measurement event viewed as a rectangle $`X_A\times Y_B`$, it is convenient to speak of $`X_A`$ and $`Y_B`$ separately as pertaining to components of the measurement event. ## III Positive operator-valued measures and conditional probabilities The diagram (13) fails for generic POVMs, but the specialization to tensor-products holds for POVMs. How this works can be seen from Neumark’s theorem, along with a corollary developed below. ### III.1 Neumark’s theorem Suppose a Hilbert space $``$ is a subspace of a Hilbert space $`^+`$. Let $`E^+:(\mathrm{\Omega })(^+)`$ be any projection-valued measure on $`^+`$, and let $`Q(^+)`$ be the orthogonal (hence self-adjoint) projection on $``$. (Although consistency calls for writing this as $`Q^+`$, to avoid clutter I write just $`Q`$.) Define a POVM $`(QE^+Q)_{}:(\mathrm{\Omega })()`$ by $$(X(\mathrm{\Omega }))(QE^+Q)_{}(X)=(QE^+(X)Q)_{},$$ (25) where $``$ as a subscript denotes the restriction of an operator to $``$. Neumark neumark ; Ak proved that all POVMs can be expressed this way: for any POVM $`M`$ on any Hilbert space $``$, there exists a Hilbert space $`^+`$ containing $``$ as a subspace, and there exists a projective resolution of the identity $`E^+:(\mathrm{\Omega })(^+)`$ such that $`M=(QE^+Q)_{}`$. ### III.2 Lifting the trace rule to $`^+`$ It is instructive to lift the trace rule to the extended Hilbert space $`^+`$. When we view $``$ as a subspace of $`^+`$, we see any vector $`|\psi `$ as a vector $`|\psi |0^{}`$ in $`^+`$, where $`|0^{}`$ denotes the 0-vector in $`^{}`$. Correspondingly, we view any operator $`A()`$ as an operator $`A\mathrm{𝟎}^{}(^+)`$, where $`\mathrm{𝟎}^{}`$ is the zero operator on $`^{}`$. With $`Q(^+)`$ the orthogonal projection onto $``$ as above, useful elementary facts are $`(A,B())(C^+,D^+(^+))`$ $`Q(A\mathrm{𝟎}^{})Q`$ $`=`$ $`A\mathrm{𝟎}^{},`$ (26) $`QC^+Q`$ $`=`$ $`(QC^+Q)_{}\mathrm{𝟎}^{},`$ (27) $`(A\mathrm{𝟎}^{})(B\mathrm{𝟎}^{})`$ $`=`$ $`(AB)\mathrm{𝟎}^{},`$ (28) $`\text{Tr}^+(A\mathrm{𝟎}^{})`$ $`=`$ $`\text{Tr}(A),`$ (29) where Tr denotes the trace on $``$ and $`\text{Tr}^+`$ denotes the trace on $`^+`$. From these equations follows Lemma: $`\text{Tr}^+[(A\mathrm{𝟎}^{})C^+]`$ $`=`$ $`\text{Tr}^+[Q(A\mathrm{𝟎}^{})QC^+]`$ (30) $`=`$ $`\text{Tr}^+[(A\mathrm{𝟎}^{})QC^+Q]`$ $`=`$ $`\text{Tr}^+\{(A\mathrm{𝟎}^{})[(QC^+Q)_{}\mathrm{𝟎}^{}]\}`$ $`=`$ $`\text{Tr}^+[A(QC^+Q)_{}\mathrm{𝟎}^{}]`$ $`=`$ $`\text{Tr}[A(QC^+Q)_{}].`$ Letting $`A`$ be $`\rho `$ and $`C^+`$ be $`E^+(X)`$ yields Proposition: $$(\rho )(X(\mathrm{\Omega }))\mathrm{Pr}(X|\rho )\text{Tr}[\rho M(X)]=\text{Tr}^+[(\rho \mathrm{𝟎}^{})E^+(X)].$$ (31) Eq. (31) allows the trace rule for any given single POVM to be lifted to $`^+`$, as illustrated in the commutative diagram E+(X) (ρ𝟎)(+)|Tr+[(ρ𝟎)E+(X)]|||Q,||=|Q,|||Tr[ρM(X)]M(X) ρ().fragmentsE(X) missing-subexpression fragments(ρdirect-sum0perpendicular-to)B(H)fragmentsmissing-subexpressionmissing-subexpressionmissing-subexpressionfragments|missing-subexpressionfragmentsTr[(ρdirect-sum0perpendicular-to)E(X)]missing-subexpression||missing-subexpressionmissing-subexpressionmissing-subexpression|fragmentsQ,||missing-subexpressionmissing-subexpressionfragments|Q,||missing-subexpressionmissing-subexpressionmissing-subexpression|missing-subexpressionfragmentsTr[ρM(X)]missing-subexpressionfragmentsM(X) missing-subexpression fragmentsρB(H){\begin{tabular}[]{ccccc}$E^{+}(X)$&\hskip-14.0pt\lower 5.0pt\hbox{\rotatebox{315.0}{$\longmapsto$}}&&\lower 1.99997pt\hbox{\rotatebox{225.0}{$\longmapsto$}}&$(\rho\oplus\bm{0}^{\perp})\in\mathcal{B}(\mathcal{H}^{+})$\\[-5.0pt] \tiny$\mskip 0.3mu-$&&&&\hskip-36.0pt\tiny$\mskip 0.3mu-$\\[-15.5pt] $|$&&$\mbox{Tr}^{+}[(\rho\oplus\bm{0}^{\perp})E^{+}(X)]$&&\hskip-36.0pt$|$\\[-16.0pt] $|$&&&&\hskip-36.0pt$|$\\[-16.0pt] $Q,|_{\mathcal{H}}\ |\hphantom{Q,|_{\mathcal{H}}\ }$&&$=$&&\hskip-36.0pt$\hphantom{Q,|_{\mathcal{H}}\ }|\ Q,|_{\mathcal{H}}$\\[-16.0pt] $|$&&&&\hskip-36.0pt$|$\\[-13.0pt] $\downarrow$&&$\mbox{Tr}[\rho M(X)]$&&\hskip-36.0pt$\downarrow$\\[0.0pt] $M(X)$&\hskip-14.0pt\raise 6.00006pt\hbox{\rotatebox{45.0}{$\longmapsto$}}&&\raise 9.49997pt\hbox{\rotatebox{135.0}{$\longmapsto$}}&\hskip-34.0pt$\rho\in\mathcal{B}(\mathcal{H})$\end{tabular}.} (32) Diagram 3: Lifting of the trace rule to $`^+`$ via Neumark’s theorem. ### III.3 Obstacle to expressing conditional probabilities with POVMs The desire to extend the diagram engendered by Proposition (31) to $`\mathrm{Pr}(XY|\rho )`$ for the set intersection $`XY`$ encounters the following obstacle: E+(X)E+(XY)=E+(X)E+(Y) E+(Y)| |||||Q,||M(XY)=[QE+(X)E+(Y)Q]|||||M(X)=[QE+(X)Q]M(X)M(Y)=[QE+(X)Q][QE+(Y)Q] M(Y)=[QE+(Y)Q].fragmentsE(X)fragmentsE(XY)E(X)E(Y) fragmentsE(Y)fragmentsmissing-subexpressionmissing-subexpressionmissing-subexpressionfragments|missing-subexpression missing-subexpression||missing-subexpressionmissing-subexpressionmissing-subexpression||missing-subexpressionmissing-subexpressionmissing-subexpression|fragmentsQ,||missing-subexpressionfragmentsM(XY)[QE(X)E(Y)Q]missing-subexpression||missing-subexpressionmissing-subexpressionmissing-subexpression||missing-subexpressionmissing-subexpressionmissing-subexpression|missing-subexpressionmissing-subexpressionfragmentsM(X)[QE(X)Q]fragmentsfragmentsM(X)M(Y)[QE(X)Q][QE(Y)Q] fragmentsM(Y)[QE(Y)Q]{\begin{tabular}[]{ccccc}$E^{+}(X)$&\hskip-6.0pt$\longmapsto$\hskip-6.0pt&$E^{+}(X\!\cap\!Y)\!=\!E^{+}(X)E^{+}(Y)$&\hskip-6.0pt\raise 5.5pt\hbox{\rotatebox{180.0}{$\longmapsto$}}\hskip-6.0pt&$E^{+}(Y)$\\[-7.0pt] \tiny$\mskip 0.3mu-$&&&&\tiny$\mskip 0.3mu-$\\[-15.5pt] $|$&&\hskip-16.0pt\raise 5.5pt\hbox{\rotatebox{270.0}{$\longmapsto$}}&&$|$\\[-16.0pt] $|$&&&&$|$\\[-16.0pt] $|$&&&&$|$\\[-13.0pt] $Q,|_{\mathcal{H}}\ |\hphantom{Q,|_{\mathcal{H}}\ }$&&$M(X\cap Y)\!=\![QE^{+}(X)E^{+}(Y)Q]_{\mathcal{H}}$&&$|$\\[-16.0pt] $|$&&&&$|$\\[-16.0pt] $|$&&&&$|$\\[-13.0pt] $\downarrow$&&$\hskip-16.0pt\neq$&&$\downarrow$\\[0.0pt] $M(X)\!=\![QE^{+}(X)Q]_{\mathcal{H}}$&$\!\longmapsto\!$&$M(X)M(Y)\!=\![QE^{+}(X)Q]_{\mathcal{H}}[QE^{+}(Y)Q]_{\mathcal{H}}$&$\!\raise 5.5pt\hbox{\rotatebox{180.0}{$\longmapsto$}}\!$&$M(Y)\!=\![QE^{+}(Y)Q]_{\mathcal{H}}$\end{tabular}.} (33) Diagram 4: Obstacle to lifting of product of POVMs. Because $$[QE^+(X)E^+(Y)Q]_{}\mathbf{}[QE^+(X)Q]_{}[QE^+(Y)Q]_{},$$ (34) we find that except in uninteresting special cases $$M(XY)M(X)M(Y).$$ (35) Correspondingly, efforts to define reduced states corresponding to a measurement modeled by $`M(X)`$ followed by a measurement modeled by $`M(Y)`$ encounter conceptual difficulties, touched on by Braunstein and Caves BC and discussed below in connection with probes. ### III.4 Tensor products of POVMs The obstacle to operator products of POVMs is no impediment to tensor products. Consider a POVM $`M_A:(\mathrm{\Omega }_A)(_A)`$ and another POVM $`M_B:(\mathrm{\Omega }_B)(_B)`$. By Neumark’s theorem, both of these POVMs can be expressed as restrictions of projections of projection-valued measures on the respective extended Hilbert spaces: $`E_A^+:(\mathrm{\Omega }_A)(_A^+)`$ and $`E_B^+:(\mathrm{\Omega }_B)(_B^+)`$, respectively. Let $`Q_A(_A^+)`$ be the orthogonal (hence self-adjoint) projection on $`_A`$, and similarly $`Q_B(_B^+)`$. Because a tensor product of projections is a projection, Diagram 2 for the extended Hilbert spaces extends downward, via $`Q_A`$ and $`Q_B`$ followed by reductions, to $``$. The neat thing here is that $$(Q_AQ_B)(E_A^+E_B^+)(Q_AQ_B)=(Q_AE_A^+Q_A)(Q_BE_B^+Q_B),$$ (36) the restriction of which to the subspace $`_A_B`$ is just $`M_AM_B`$; i.e. we have $`M_AM_B`$ $`=`$ $`[Q_AE_AQ_AQ_BE_BQ_B]|_{_{BA}}`$ (37) $`=`$ $`[(Q_AQ_B)(E_AE_B)(Q_AQ_B)]|_{_{BA}}.`$ Thus we arrive at the diagram: $$\begin{array}{ccccccccccc}& E^+_A& & E^+_A\mathrm{𝟏}_B& & E^+_AE^+_B& & \mathrm{𝟏}_AE_B^+& & E^+_B& \\ Q_A,|__A& & & & & & & & & & Q_B,|__B\\ & M_A& & M_A\mathrm{𝟏}_B& & M_AM_B& & \mathrm{𝟏}_AM_B& & M_B& \end{array}.$$ (38) Diagram 5: Tensor product of POVMs. Because this shows $$(M_AM_B)[(X_A\times \mathrm{\Omega }_B)(\mathrm{\Omega }_A\times Y_B)]=M_A(X_A)M_B(Y_B),$$ (39) Eqs. (19)–(23) hold also for non-projective POVMs; for instance, Eq. (23) becomes (with the understanding $`XX_A`$ and $`YY_B`$): $`\mathrm{Pr}(X|Y)`$ $`=`$ $`{\displaystyle \frac{\text{Tr}_{AB}[M_A(X)M_B(Y)]\rho _{AB}}{\text{Tr}_{AB}\{\rho _{AB}[\mathrm{𝟏}_AM_B(Y)]\}}}`$ (40) $`=`$ $`{\displaystyle \frac{\text{Tr}_A\text{Tr}_B\{[\mathrm{𝟏}_AM_B(Y)]\rho _{AB}[\mathrm{𝟏}_AM_B(Y)](M_A(X)\mathrm{𝟏}_B)\}}{\text{Tr}_{AB}\{\rho _{AB}[\mathrm{𝟏}_AM_B(Y)]\}}}`$ $`=`$ $`\text{Tr}_A[\rho _A|_YM_A(X)],`$ with $`\rho _A|_Y`$ $`\stackrel{\mathrm{def}}{=}`$ $`{\displaystyle \frac{\text{Tr}_B\{[\mathrm{𝟏}_AM_B(Y)]\rho _{AB}[\mathrm{𝟏}_AM_B(Y)]\}}{\text{Tr}_{AB}\{\rho _{AB}[\mathrm{𝟏}_AM_B(Y)]\}}}={\displaystyle \frac{\text{Tr}_B\{\rho _{AB}[\mathrm{𝟏}_AM_B(Y)]\}}{\text{Tr}_{AB}\{\rho _{AB}[\mathrm{𝟏}_AM_B(Y)]\}}}.`$ (41) ### III.5 Tensor products of families of POVMs Instead of considering a single POVM on $`_A`$ and a single POVM on $`_B`$, we consider now two indexed families of POVMs, $`M_{A,\alpha }`$ and $`M_{B,\beta }`$. Then Diagram 5 threatens to become decorated with these indices in a curiously asymmetric way, because the proof of Neumark’s theorem presents $`_B`$ not as a $`\beta `$-independent subspace of $`_B^+`$, but as one that varies with $`\beta `$. The following corollary restores symmetry to the distribution of indices and justifies lifting not just one POVM but any family of POVMs to a single extended Hilbert space related to the base space by a single (index-independent) orthogonal projection. Corollary to Neumark’s theorem: Given a fixed Hilbert space $``$ and a fixed $`\sigma `$-algebra $`(\mathrm{\Omega })`$ of measurable sets, together with any family of POVMs $`M_\beta :(\mathrm{\Omega })()`$ (indexed by $`\beta `$); then there is a single extended Hilbert space $`^+`$ containing $``$ as a subspace and a single $`Q(^+`$), the orthogonal projection on $``$, along with a family of projection-valued measures $`E_\beta ^+`$ on $`^+`$, such that $$(\beta )(E_\beta ^+)M_\beta =[QE_\beta ^+Q]_{}.$$ (42) Proof sketch: The proof of Neumark’s theorem in Ak (and also in neumark ) deals with each POVM of a family one at a time, so to speak: it constructs an extended Hilbert space $`^+`$ that depends on $`(\mathrm{\Omega })`$ but is independent of $`\beta `$; then $``$ is mapped isomorphically onto a subspace that I denote $`\stackrel{~}{}_\beta ^+`$ that depends on $`\beta `$; implicit in the proof is an isomorphism carrying each density operator $`\rho ()`$ to $`\stackrel{~}{\rho }_\beta (\stackrel{~}{}_\beta )`$; this works in such a way that $`E^+`$ is independent of $`\beta `$. We want to swap these dependencies, to make $`\stackrel{~}{}`$ independent of $`\beta `$ in exchange for allowing a $`\beta `$-dependent $`E_\beta ^+`$. I claim this works as follows. Pick any value of $`\beta `$, and call it 0; for this value of $`\beta `$, $``$ is mapped isomorphically onto $`\stackrel{~}{}_0`$. For any value of $`\beta `$ there exists a unitary transform $`U_\beta ^+(^+)`$, such that $`\stackrel{~}{\rho }_\beta \stackrel{~}{\mathrm{𝟎}}_\beta ^{}=U_\beta ^+(\rho \mathrm{𝟎}^{})U_\beta ^{}.`$ From this and Proposition (31) it follows that $`(\beta ,X)\text{Tr}[\rho M_\beta (X)]`$ $`=`$ $`\text{Tr}^+[E^+(X)(\stackrel{~}{\rho }_\beta +\stackrel{~}{\mathrm{𝟎}}_\beta ^{})]`$ (43) $`=`$ $`\text{Tr}^+[U_\beta ^+(\rho +\mathrm{𝟎}^{})U_\beta ^+E^+(X)]`$ $`=`$ $`\text{Tr}^+[(\rho +\mathrm{𝟎}^{})U_\beta ^+E^+(X)U_\beta ^+]`$ $`=`$ $`\text{Tr}^+[(\rho +\mathrm{𝟎}^{})E_\beta ^+(X)],`$ where we define $`E_\beta ^+(X)\stackrel{\mathrm{def}}{=}U_\beta ^+E^+(X)U_\beta ^+`$. Equation (43) and Lemma (30) imply that for an orthogonal projection $`Q`$ onto $`\stackrel{~}{}`$, independent of $`\beta `$, $`(\beta ,X,\rho )\text{Tr}[\rho M_\beta (X)]=\text{Tr}[\rho (QE_\beta ^+(X)Q)_\stackrel{~}{}],`$ from which the conclusion of the corollary follows. $`\mathrm{}`$ A simple example in finite dimensions occurs in MB97 , where a plane on which a POVM is defined is embedded in a $`\beta `$-dependent way in a three space equipped with a fixed projection-valued measure defined by three mutually orthogonal basis vectors. As in the corollary, one can just as well hold the plane fixed and rotate the basis vectors. We apply this as follows. Consider a family of POVMs indexed by $`\alpha `$, $`M_{A,\alpha }:(\mathrm{\Omega }_A)(_A)`$, along with another family of POVMs indexed by $`\beta `$, $`M_{B,\beta }:(\mathrm{\Omega }_B)(_B)`$. By the corollary, we translate Diagram 5 into a diagram for these families. For all $`\alpha `$, $`\beta `$ we have: $$\begin{array}{ccccccccccc}& E^+_{A,\alpha }& & E^+_{A,\alpha }\mathrm{𝟏}_B& & E^+_{A,\alpha }E^+_{B,\alpha }& & \mathrm{𝟏}_AE^+_{B,\beta }& & E^+_{B,\alpha }& \\ Q_A,|__A& & & & & & & & & & Q_B,|__B\\ & M_{A,\alpha }& & M_{A,\alpha }\mathrm{𝟏}_B& & M_{A,\alpha }M_{B,\beta }& & \mathrm{𝟏}_AM_{B,\beta }& & M_{B,\beta }& \end{array}.$$ (44) Diagram 6: Tensor product of POVMs of two families. By the corollary, we have a diagram in which $`Q_A`$ and $`Q_B`$ are independent of the indices $`\alpha `$, $`\beta `$. ## IV Applying quantum probabilities to descriptions of devices Designs for systems of devices often start with models expressed in simplified equations; quantum cryptography is a case in point. Implementing a design inspired by equations entails arranging devices—lasers, detectors, counters—so that measured device behavior accords with properties expressed in the equations. Experience teaches that the devices work as desired only when nudged in ways unexpressed by the starting equations. To support this nudging, one ends up with layers of more detailed equations, needed for instance in order to design feedback loops that compensate for various drifts. Whatever details we undertake to model, we face choices. For example, in an experiment with pulsed light, if we assume there are no memory effects in the light detectors, we can implement the state preparation for a trial by generating a single light pulse, while to study memory effects in detectors, we must implement the state preparation for a trial by generating a sequence of pulses ams . When the equations used in modeling are equations in the language of quantum mechanics, different choices are expressed by different probability distributions stemming from different density operators and different positive operator-valued measures (POVMs), along with different choices of a measure space. ### IV.1 Choice of measure space To use a measure space as part of a quantum model of devices, one needs, somehow, to link parameters—frequencies, positions, times—by which one speaks in the laboratory of light pulses to the measure space. Several things complicate this linking. First, no single-frequency light state is an element of any separable Hilbert space. Second are laboratory facts (filters spill over, light diffracts, signal times are fuzzy). In modeling these facts, we need detection operators that are correspondingly unsharp in relation to the same parameters. Third, the specification of a measure space in terms of physical parameters depends on the layer of detail, and this impedes comparisons of models across different layers of detail. These complications are eased by adapting a trick from quantum decision theory, in which POVMs are defined not with respect to an arbitrary measure space but only for a discrete measure space, the elements of which are thought of not in terms of physical parameters but as possible actions to take in response to a measurement event. So far the measure spaces appearing in this report have been arbitrary, i.e., without any additional constraints. They allow for real metric spaces needed to deal with continuous spectra of hermitian detection operators, and these measure spaces necessarily involve the physical parameters of the spectrum. We can preserve the capacity to deal with these parameters when we must, while sidestepping their complications when we can: the trick is to link arbitrary measure spaces to discrete measure spaces adapted from quantum decision theory helstrom ; peres . In quantum decision theory, a POVM is defined not with reference to a general measure space but as a countable set of detection operators $`M(j)`$ that sum to $`\mathrm{𝟏}`$. One can imagine the measurement result displayed by lighting up just one of a row of lamps, indexed by $`j`$. The context is one of action: “if light $`j`$ goes on, do $`X_j`$.” To link this to a general measure space $`\mathrm{\Omega }`$, we model an act of measuring not by a full POVM, but by a countable set of detection operators $`M(\mathrm{\Omega }_j)`$, where for $`j=1,2,\mathrm{},`$ the $`\mathrm{\Omega }_j`$ are mutually disjoint subsets that cover $`\mathrm{\Omega }`$. When we have in mind a mapping from natural numbers to measurable sets of some measure space $`\mathrm{\Omega }`$, we can abbreviate $`M(\mathrm{\Omega }_j)`$ by $`M(j)`$. This abbreviation establishes a relation among detection operators at different levels of detail with distinct measure spaces $`\mathrm{\Omega }`$ and $`\mathrm{\Omega }^{}`$ by relating say $`M(\mathrm{\Omega }_j)`$ and $`M(\mathrm{\Omega }_j^{})`$ to the same decision-oriented event $`j`$. By this linking of any arbitrary measure space to the natural numbers as a single discrete measure space, we make universally applicable the notion of a family of POVMs defined on that discrete measure space. ## V Two applications of conditional probabilities Here are two applications of conditional probabilities as discussed above, the first conceptual, the second concrete. ### V.1 Sequences of probes in place of “consecutive measurements” Tensor-product spaces are well suited to the modeling of the interaction of a particle with a succession of probes, followed by measurement of the probes, and in this connection a variety of conditional density operators can be useful. For example, express the particle to be probed by a density operator $`\rho _0(_0)`$, and express probe $`j`$, $`j=1,2,\mathrm{},`$ prior to its interaction with the particle by a density operator $`\rho _j(_j)`$. After a succession of interactions (shown in Fig. 1), expressed by unitary operators $`U_{0j}(_0_j)`$, the probes are measured, as expressed by POVMs $`M_j:(\mathrm{\Omega }_j)(_j)`$ on the respective Hilbert spaces $`_j`$. For two probes, this procedure yields for the joint probability of components $`X_j`$ of the measurement event: $`\mathrm{Pr}(X_1,X_2)`$ $`=`$ $`\text{Tr}_{012}\left(M_2(X_2)M_1(X_1)U_{02}\{[U_{01}(\rho _0\rho _1)U_{01}^{}]\rho _2\}U_{02}^{}\right)`$ (45) $`=`$ $`\text{Tr}_{02}[M_2(X_2)U_{02}(\sigma _0|_{X_1}\rho _2)U_{02}^{}],`$ where $$\sigma _0|_{X_1}=\text{Tr}_1[M(X_1)U_{01}(\rho _0\rho _1)U_{01}^{}].$$ (46) The corresponding conditional probability of $`X_2`$ given $`X_1`$ is $$\mathrm{Pr}(X_2|X_1)=\text{Tr}_{02}[M_2(X_2)U_{02}(\rho _0|_{X_1}\rho _2)U_{02}^{}],$$ (47) where $`\rho _0|_{X_1}`$ is the (normalized) conditional density operator engendered by the scaled conditional density operator $`\sigma _0|_{X_1}`$: $$\rho _0|_{X_1}\stackrel{\mathrm{def}}{=}\frac{\sigma _0|_{X_1}}{\text{Tr}_{02}\sigma _0|_{X_1}}.$$ (48) For three probes, this procedure yields for the joint probability of components $`X_j`$ of the measurement event: $`\mathrm{Pr}(X_1,X_2,X_3)`$ $`=`$ $`\text{Tr}_{023}\left[M_3(X_3)M_2(X_2)U_{03}\left([U_{02}(\sigma _0|_{X_1}\rho _2)U_{02}^{}]\rho _3\right)U_{03}^{}\right]`$ (49) $`=`$ $`\text{Tr}_{03}[M_3(X_3)U_{03}(\sigma _0|_{X_1,X_2}\rho _3)U_{03}^{}],`$ where $`\sigma _0|_{X_1}`$ is given in Eq. (46) and $`\sigma _0|_{X_1,X_2}`$ $`=`$ $`\text{Tr}_2[M(X_2)U_{02}(\sigma _0|_{X_1}\rho _2)U_{02}^{}].`$ (50) Corresponding to these equations, we find for the conditional probabilities $`\mathrm{Pr}(X_2,X_3|X_1)`$ $`=`$ $`\text{Tr}_{023}\left[M_3(X_3)M_2(X_2)U_{03}\left([U_{02}(\rho _0|_{X_1}\rho _2)U_{02}^{}]\rho _3\right)U_{03}^{}\right],`$ (51) $`\mathrm{Pr}(X_3|X_1,X_2)`$ $`=`$ $`\text{Tr}_{03}[M_3(X_3)U_{03}(\rho _0|_{X_1,X_2}\rho _3)U_{03}^{}],`$ (52) where each conditional density operator is defined as usual by dividing a corresponding scaled conditional density operator by its trace. Recognizing freedom to invoke probes such as these in modeling measurements has been shown to make the notion of repeated measurements unnecessary in formulating the mathematics of quantum mechanics in terms of projection-valued measures: a pair of successive measurements can be subsumed into a single measurement involving a succession of interactions of probes 0404113 . By virtue of Neumark’s theorem and the corollary above which makes it applicable to a family of POVMs, this now generalizes to POVMs. Once freedom to invoke probes is accepted, there is no place nor any need in the logic of quantum mechanics for a postulate pertaining to consecutive measurements. In particular, notwithstanding the efforts of Dirac dirac and von Neumann vN and others luders ; davies concerning consecutive measurements, there is no place for a postulate of so-called “state reductions.” ### V.2 Example from quantum cryptography Models of quantum key distribution (QKD) subject to individual eavesdropping attacks qkd ; slutsky posit a sequence of trials, one trial for each raw key bit. The popular key protocol BB84 BB84 has two versions, “transmitted-state” (Fig. 2a) and “entangled-state” (Fig. 2b). Transmitted-state BB84 calls at each trial for Alice to prepare at random one of four light states $`\rho _B(i)`$, $`i=1,\mathrm{},4,`$ with prior probabilities $`\zeta _i`$; these states, subject to probing by Eve, are detected by Bob. Entangled-state BB84 calls for Alice to prepare a single polarization-entangled light state $`\rho _{BA}`$ that propagates to both her own detectors and to Bob’s detectors, again with the propagation to Bob subject to probing by Eve. Here I develop a relation that enables one to formulate individual attacks against the seemingly more complicated entangled-state BB84 in the same way as individual attacks against transmitted-state BB84. The formulation by Slutsky et alslutsky for transmitted-state BB84 holds for general light states defined in jm05 ; Frwk , not just the simplified states for which they carried through their analysis. As discussed in detail in jm05 ; SPIE , all the probabilities pertinent to key distribution in the face of an individual eavesdropping attack against transmitted-state BB84 stem from $`\zeta _i`$ together with the trace rule applied to a quantum state on a tensor-product space $`_E_B`$, where $`_B`$ is a Hilbert space of light modes transmitted to Bob and $`_E`$ is the Hilbert space of Eve’s probe. These probabilities have the form $$\mathrm{Pr}(X_E,Y_B,\rho _B(i))=\zeta _i\text{Tr}_{EB}[M_E(X_E)M_B(Y_B)U_{EB}(\rho _E\rho _B)U_{EB}^{}],$$ (53) where $`U_{EB}`$ represents an arbitrary unitary interaction chosen by Eve between her probe $`\rho _E`$ and Bob’s light state $`\rho _B`$. In models of entangled-state BB84, there are no prior probabilities $`\zeta _i`$; instead, the Hilbert space involves three factors $`_E_B_A`$, where $`_A`$ is the Hilbert space for light detected by Alice. At each trial Eve prepares a probe $`\rho _E`$ as before, but now Alice prepares an entangled state $`\rho _{BA}(_B_A)`$. With $`U_{EB}`$ as in transmitted-state BB84, the probabilities pertinent to key distribution in the face of an individual eavesdropping attack against entangled-state BB84 are just $$\mathrm{Pr}(X_E,Y_B,Z_A)=\text{Tr}_{EBA}[M_E(X_E)M_B(Y_B)M_A(Z_A)U_{EB}(\rho _E\rho _{BA})U_{EB}^{}],$$ (54) which looks significantly different from Eq. (53); however, by the partial trace manipulations of Appendix A, in particular Eq. (79), this last equation becomes $$\mathrm{Pr}(X_E,Y_B,Z_A)=\text{Tr}_{EB}\{M_E(X_E)M_B(Y_B)U_{EB}[\rho _E\text{Tr}_A(M_A(Z_A)\rho _{BA})]U_{EB}^{}\}.$$ (55) This implies the following probability for Alice’s component $`Z_A`$ of a measurement event (regardless of Bob’s and Eve’s): $`\mathrm{Pr}(Z_A)`$ $`=`$ $`\text{Tr}_{EB}\{U_{EB}[\rho _E\text{Tr}_A(M_A(Z_A)\rho _{BA})]U_{EB}^{}\}`$ (56) $`=`$ $`\text{Tr}_{EB}[\rho _E\text{Tr}_A(M_A(Z_A)\rho _{BA})]`$ $`=`$ $`\text{Tr}_B[\text{Tr}_A(M_A(Z_A)\rho _{BA})],`$ which takes the part of $`\zeta _i`$ in transmitted-state BB84. Taking the part of Eq. (53) is the conditional probability $`\mathrm{Pr}(X_E,Y_B|Z_A)`$ $`\stackrel{\mathrm{def}}{=}`$ $`{\displaystyle \frac{\mathrm{Pr}(X_E,Y_B,Z_A)}{\mathrm{Pr}(Z_A)}}`$ (57) $`=`$ $`{\displaystyle \frac{\text{Tr}_{EB}\{M_E(X_E)M_B(Y_B)U_{EB}[\rho _E\text{Tr}_A(M_A(Z_A)\rho _{BA})]U_{EB}^{}\}}{\text{Tr}_B[\text{Tr}_A(M_A(Z_A)\rho _{BA})]}}`$ $`=`$ $`\text{Tr}_{EB}[M_E(X_E)M_B(Y_B)U_{EB}(\rho _E\rho _B|_{Z_A})U_{EB}^{}],`$ with the conditional density operator $$\rho _B|_{Z_A}\stackrel{\mathrm{def}}{=}\frac{\text{Tr}_A[M_A(Z_A)\rho _{BA}]}{\text{Tr}_B[\text{Tr}_A(M_A(Z_A)\rho _{BA})]}.$$ (58) Entangled-state BB84 requires that the state-event components $`Z_A`$ for Alice include four that correspond to the four choices Alice has in transmitted-state BB84. I call these $`Z_{A,i}`$. By virtue of Eqs. (56) and (57) we arrive at the following Proposition: Any model of entangled-state BB84 of the form of Eq. (54) asserts the same joint probabilities relevant to quantum key distribution subject to individual eavesdropping attacks as does the model of transmitted-state BB84 with $`\zeta _i=\mathrm{Pr}(Z_{A,i})`$ defined by Eq. (56) and with $`\rho _B(i)=\rho _B|_{Z_{A,i}}`$ defined by Eq. (58). ###### Acknowledgements. For very helpful discussions over the past four years, I thank Tai Tsun Wu. This work was supported in part by the Air Force Research Laboratory and DARPA under Contract No. F30602-01-C-0170 with BBN Technologies. ## Appendix A Tensor products and partial traces By a bounded positive operator $`A`$ on a Hilbert space $``$, I mean any bounded self-adjoint operator such that $`(|x)x|A|x0`$. The trace of a positive operator $`A`$ on a separable Hilbert space is defined sewell as $$\text{Tr}(A)=\underset{n}{}\psi _n|A|\psi _n,$$ (59) where $`\{\psi _n\}`$ is an orthonormal basis. In fact, the value of $`\text{Tr}(A)`$ here, which might be infinite, is independent of the choice of basis vN . As stated in sewell and proved in vN , the trace is invariant under cyclic permutations of the factors of a product: $$\text{Tr}(ABC)=\text{Tr}(CAB).$$ (60) For a finite-dimensional vector space, the following discussion is elementary; a start at the more complicated derivations needed for infinite-dimensional, separable Hilbert spaces can be found in Chap. II, Sec. 11 of vN . The trace of a product of matrices acting on a finite-dimensional vector space $``$ is defined by $$\text{Tr}(MN)=\underset{j,k}{}M_{j,k}N_{k,j}=\underset{j,k}{}N_{j,k}M_{k,j}=\text{Tr}(NM).$$ (61) Suppose that $`=_A_B`$. For the tensor product of two vectors $`a_A`$ and $`b_B`$, with components $`a_J`$ and $`b_j`$, respectively, we write the tensor product as a vector $`w_A_B`$ with components $`w_{Jj}=a_Jb_j`$. Now let $`M^{(AB)}`$ be a matrix operating on vectors of $`_A_B`$. A row of such a matrix is specified by a double index such as $`Jj`$, and a column by a double index such as $`Kk`$, so that in the special case $`M^{(AB)}=R^{(A)}S^{(B)}`$, $`M_{JjKk}^{(AB)}=R_{JK}^{(A)}S_{jk}^{(B)}`$ (where we have assumed $`R^{(A)}`$ is a matrix acting on vectors of $`_A`$ and $`S^{(B)}`$ is a matrix acting on vectors of $`_B`$). The partial trace over the $`_B`$ factor of $`M^{(AB)}`$ is defined to be the matrix acting on $`_A`$ that has as its $`(J,K)`$-th component $$[\text{Tr}_B(M^{(AB)})]_{JK}=\underset{j}{}M_{JjKj}.$$ (62) For the full trace over $`_A_B`$ I write $`\text{Tr}_{AB}`$. Any matrix $`M^{(AB)}`$ can be written as a sum of tensor products, with the result that the properties of partial traces follow from block form of these tensor-product terms. Recall that for finite-dimensional spaces the tensor product of any two matrices $`A`$ and $`B`$ is a matrix of $`B`$-sized blocks: $$AB=\left[\begin{array}{ccc}A_{11}B\hfill & A_{12}B\hfill & \mathrm{}\hfill \\ A_{21}B\hfill & A_{22}B\hfill & \mathrm{}\hfill \\ \mathrm{}\hfill & \mathrm{}\hfill & \mathrm{}\hfill \end{array}\right],$$ (63) from which it is obvious that for square matrices $`A`$ and $`B`$, $$\text{Tr}_B(AB)=\left[\begin{array}{ccc}A_{11}\text{Tr}(B)\hfill & A_{12}\text{Tr}(B)\hfill & \mathrm{}\hfill \\ A_{21}\text{Tr}(B)\hfill & A_{22}\text{Tr}(B)\hfill & \mathrm{}\hfill \\ \mathrm{}\hfill & \mathrm{}\hfill & \mathrm{}\hfill \end{array}\right]=[\text{Tr}(B)]A.$$ (64) Similarly, we have $`\text{Tr}_A(AB)=[\text{Tr}(A)]B`$, from which it follows immediately that $`\text{Tr}_{AB}(AB)=[\text{Tr}_A(A)][\text{Tr}_B(B)]`$. Except when computing with matrix components, I drop the superscripts indicating “which space” to subscripts, writing $`M_{AB}`$ in place of $`M^{(AB)}`$, etc. Although the full trace of a product is invariant under a change in the order of the factors, the block form makes apparent that changing the order of factors affects the partial trace of their product; in the general case, $`\text{Tr}_B[(R_AS_B)(R_A^{}S_B^{})]`$ $`=`$ $`\text{Tr}_B[(R_AR_A^{})(S_BS_B^{})]=[\text{Tr}(S_BS_B^{})]R_AR_A^{}`$ (65) $`\mathbf{}`$ $`\text{Tr}_B[(R_A^{}S_B^{})(R_AS_B)].`$ It is of course true that for any two matrices $`M`$ and $`N`$ acting on $`_A_B`$, we have $`(M,N)\text{Tr}_A[\text{Tr}_B(MN)]`$ $`=`$ $`\text{Tr}_A[\text{Tr}_B(NM)]`$ (66) $`=`$ $`\text{Tr}_B[\text{Tr}_A(MN)]=\text{Tr}_B[\text{Tr}_A(NM)]=\text{Tr}_{AB}(MN).`$ Easily proved are the following facts concerning square matrices, stated with subscripts to indicate the relevant vector spaces, $`_A`$, $`_B`$, and $`_A_B`$. First, there holds a commutativity relation for partial traces of factors of a certain form: Lemma: $$(S_B,M_{AB})\text{Tr}_B[M_{AB}(\mathrm{𝟏}_AS_B)]=\text{Tr}_B[(\mathrm{𝟏}_AS_B)M_{AB}].$$ (67) Proof: Expand any $`M_{AB}`$ as a sum of tensor products $`R_AS_B`$, and observe that for this special case the $``$ of Eq. (65) becomes equality. $`\mathrm{}`$ The same technique shows Lemma: $$(\text{ hermitian }S_B,\text{ hermitian }M_{AB})\text{Tr}_B[M_{AB}(\mathrm{𝟏}_AS_B)]\text{ is hermitian.}$$ (68) Again the same technique shows Lemma: $`\text{Tr}_B[(R_A\mathrm{𝟏}_B)M_{AB}]`$ $`=`$ $`R_A\text{Tr}_B(M_{AB}),`$ (69) $`\text{Tr}_B[M_{AB}(R_A\mathrm{𝟏}_B)]`$ $`=`$ $`[\text{Tr}_B(M_{AB})]R_A,`$ (70) from which follows Lemma: $$(R_A,M_{AB})\text{Tr}_{AB}[(R_A\mathrm{𝟏}_B)M_{AB}]=\text{Tr}_A[R_A\text{Tr}_B(M_{AB})].$$ (71) Note the stretch of notation so that $`\text{Tr}_A`$ is used here for the trace on $`_A`$, while it is also used, as above, for the partial trace over the $`A`$ factor of $`_A_B`$. All these lemmas holds when $`A`$ and $`B`$ are interchanged; for instance the last lemma becomes: $$(S_B,M_{AB})\text{Tr}_{AB}[(\mathrm{𝟏}_AS_B)M_{AB}]=\text{Tr}_B[S_B\text{Tr}_A(M_{AB})].$$ (72) Here is an example of the application to quantum mechanics. Sometimes one wants to compute a conditional density operator of the form $`\rho _A`$ $`=`$ $`\text{Tr}_A\left(M_A(X)|\psi _{AB}\psi _{AB}|\right)\text{Tr}_A\left((M_A(X)\mathrm{𝟏}_B)|\psi _{AB}\psi _{AB}|\right),`$ (73) where $`M_A(X)`$ is positive hermitian and thus can be written as a product of its positive square roots $`[M_A(X)]^{1/2}`$. In some cases the calculation is made easier by putting Eq. (73) in a symmetric form. To this end, interchange $`A`$ and $`B`$ in Lemma (67) and change names to obtain Lemma: $$(R_A,Q_{AB})\text{Tr}_A[(R_A\mathrm{𝟏}_B)Q_{AB}]=\text{Tr}_A[Q_{AB}(R_A\mathrm{𝟏}_B)].$$ (74) Apply this with $`R_A`$ taken as $`[M_A(X)]^{1/2}`$ and with $`Q_{AB}`$ taken as $`[M_A(X)]^{1/2}|\psi _{AB}\psi _{AB}|`$ to obtain the symmetrized form $`\rho _A`$ $`=`$ $`\text{Tr}_A\left([M_A(X)]^{1/2}|\psi _{AB}\psi _{AB}|[M_A(X)]^{1/2}\right).`$ (75) To deal with partial traces of more complex expressions, a couple of tricks help. The first is a shorthand notation. So far I have written out lots of tensor products of operators in which the operator on one of the factor spaces is the identity operator for that space, as in $`\text{Tr}_B[M_{AB}(\mathrm{𝟏}_AS_B)]`$, which contains an operator product of the operator $`M_{AB}`$ on $`_A_B`$ times the indicated tensor product $`\mathrm{𝟏}_AS_B`$. Common in the physics literature and often convenient is a shorthand convention of writing just $`\text{Tr}_B(M_{AB}S_B)`$. This shorthand can be undone by tensoring each operator in such an expression into the identity operators required for the whole expression to make sense. The second trick is merely to recognize that if $`_A`$ and $`_B`$ are distinct factor spaces of a tensor-product space $`_{AB}`$, then $$[M_A,N_B]=[(M_A\mathrm{𝟏}_B)(\mathrm{𝟏}_AN_B)]=0;$$ (76) that is, operators on distinct factor spaces commute with one another. For example, Lemmas (67) and (68) with a swapping of $`A`$ and $`B`$ become in shorthand notation $`\text{Tr}_A(M_{AB}R_A)`$ $`=`$ $`\text{Tr}_A(R_AM_{AB})`$ (77) $`\text{Tr}_A(S_BM_{AB})`$ $`=`$ $`S_B\text{Tr}_A(M_{AB}).`$ (78) From these equations follows a more complicated relation for operators on a triple tensor product $`_E_B_A`$, of use in quantum cryptography: $`\text{Tr}_A[M_EM_BM_AU_{EB}\rho _E\rho _{AB}U_{EB}^{}]`$ $`=`$ $`\text{Tr}_A[M_EM_BU_{EB}\rho _EM_A\rho _{AB}U_{EB}^{}]`$ (79) $`=`$ $`M_EM_BU_{EB}\rho _E[\text{Tr}_A(M_A\rho _{AB})]U_{EB}^{}.`$
warning/0506/astro-ph0506101.html
ar5iv
text
# On the Afterglow and Host Galaxy of GRB 021004: A Comprehensive Study with the Hubble Space Telescope1footnote 11footnote 1Based on observations made with the NASA/ESA Hubble Space Telescope, obtained at the Space Telescope Science Institute, which is operated by the Association of Universities for Research in Astronomy, Inc., under NASA contract NAS 5-26555. These observations are associated with programs 9074 and 9405. ## 1 Introduction The very rapid localization of cosmic Gamma-Ray Bursts (GRBs) made possible by the HETE-2 satellite has enabled very well-sampled multi-band lightcurves ranging from a few minutes to several months after the bursts. One of the best studied GRBs so far is the bright burst detected on 2002 October 4 (all epochs are given as UT) with HETE-2 (Shirasaki et al. 2002). The optical afterglow was detected unusually early – 3.2 minutes after the high energy event (Fox et al. 2003) and monitored intensively with many telescopes in the following hours to months (e.g., Pandey et al. 2003; Holland et al. 2003 – hereafter H03; Bersier et al. 2003; Mirabal et al. 2003 – hereafter M03). The degree of polarization was measured on several epochs leading to the detection of a variable polarization angle (Rol et al. 2003). The redshift was determined to be $`z=2.33`$ based on strong hydrogen and metal absorption lines as well as a Ly$`\alpha `$ emission line from the underlying host galaxy (Chornock & Filippenko 2002; Møller et al. 2002; Pandey et al. 2003; Matheson et al. 2003; M03). The absorption system associated with GRB 021004 contained several components covering a velocity range of more than 3000 km s<sup>-1</sup> with clear evidence for line-locking between the components (Savaglio et al. 2002; Møller et al. 2002; Fiore et al. 2004) Despite the intensive coverage of this afterglow there is quite limited agreement between the reported afterglow parameters and their interpretation, i.e., the spectral and late-time decay slopes, wind or ISM circumburst medium, and position of the cooling break (e.g., Lazzati et al. 2002; Pandey et al. 2003; Li & Chevalier 2003; Holland et al. 2003; Heyl & Perna 2003; Dado et al. 2003; Mirabal et al. 2003; Rol et al. 2003; Nakar et al. 2003; Björnsson, Gudmundsson & Jóhannesson 2004). The reasons are i) the very complex lightcurve during the first week and ii) the presence of a relatively bright host galaxy affecting the analysis of the late (fainter) part of the afterglow lightcurve. In this paper we present an analysis of HST observations of the afterglow of GRB 021004 ranging from three days to ten months after the burst. Our emphasis is on the early UV spectroscopy, the late-time afterglow and on the properties of the host galaxy. In Sect. 2 we describe the observations, data reduction, and analysis and in Sect. 3 we present our discussion and conclusions. We assume $`H_0`$ = 70 km s<sup>-1</sup> Mpc<sup>-1</sup>, $`\mathrm{\Omega }_m`$ = 0.3 and $`\mathrm{\Omega }_\mathrm{\Lambda }`$ = 0.7 throughout. ## 2 Observations and Data Reduction HST observed GRB 021004 after a fast turnaround on 2002 October 6–7. On 2002 October 6 we used the Space Telescope Imaging Spectrograph (STIS) to obtain near-UV spectroscopic observations of the afterglow. The Prism (4 orbits) as well as the G430L grating (2 orbits) were used for the spectroscopic STIS observations. The G430L spectra were obtained with a 0.5 arcsec slit and a position angle of 113.8<sup>o</sup>. The observing log is given in Table 1. The choice of the Prism could appear surprising at first sight as it provides little coverage redwards of the Lyman edge at the GRB redshift. However, at the time of submitting the Phase II proposal, the redshift was expected to be $`1.60<z<2.1`$ based on the presence of a $`z=1.60`$ Mg II system and the lack of strong Ly$`\alpha `$ absorption (Fox et al. 2002; Eracleous et al. 2002; Weidinger et al. 2002). The correct redshift $`z=2.33`$ was only announced on 2002 October 8 (Chornock & Filippenko 2002) after the first HST observations. The reduction and analysis of the spectroscopic observations are described in Sect. 2.1. Following the STIS observation the afterglow was observed with the Near Infrared Camera and Multi-Object Spectrometer (NICMOS) in the F110W and F160W filters and with the Advanced Camera for Surveys (ACS) High Resolution and Wide Field Cameras (HRC and WFC) in the four filters F250W, F435W, F606W, and F814W spanning nearly four octaves in wavelength from the near UV to the near-IR. The field was observed again on 2002 October 11, 2002 October 22, 2002 November 26, 2003 May 31, 2003 July 21, and 2003 July 26 in various WFC and NICMOS filters. The full journal of observations is listed in Table 1. The reduction and analysis of the NICMOS and ACS data are described in Sect. 2.2. ### 2.1 Spectroscopic STIS Observations Reduction of the STIS observations were performed using the STIS Instrument Definition Team version of CALSTIS (Lindler 2003). #### 2.1.1 The Prism Spectra The Prism dispersion is strongly wavelength dependent (for further details see the HST Instrument Handbook, Kim Quijano et al. 2003, and Smette et al. 2001). The dispersion is 40 Å/pixel at 3000 Å, the expected wavelength of the Lyman edge at the GRB redshift and is smaller at longer wavelengths. Consequently, a Prism spectrum is quite sensitive to errors in the flat–field, especially for $`\lambda >3000`$ Å. Also, the flux calibration is strongly dependent on the accuracy of the zero-point of the wavelength calibration. The overall calibration is unreliable for $`\lambda >3300`$ Å. Over the 300 Å useful range of our spectra, these systematic effects were reduced by dithering the object along the slit between the four exposures. In addition, the zero-point of the wavelength calibration also depends on the precise location of the object within the slit. Therefore, we compared the location of the afterglow in the acquisition image with the location of the slit center. The latter was measured by fitting a Gaussian at each row of an image of the slit on the CCD obtained with a Tungsten lamp obtained at the end of the first orbit. The centers of each fitted Gaussian were themselves fitted as a function of column number with a $`3^{\mathrm{rd}}`$ degree polynomial. At the row where the afterglow image reaches a maximum intensity, the difference between the afterglow location and the slit center gives an offset of $`0.064`$ pixels or $`0\stackrel{}{\mathrm{.}}0016`$. However, an important problem is that HST+STIS shows some flexure for the first few orbits after acquiring a new target as the telescope has changed its orientation relative to the sun. Consequently, one cannot be sure that the object stayed over the same pixel during the 4 orbits. Therefore, the possibility exists that the zero point of the wavelength calibration estimated during the first orbit may not be correct in the following ones. In order to test and correct for any change in this quantity, we consider three different ways. We checked the wavelength of the telluric Ly-alpha emission. However, the width of the line in pixels is large because of the large slit width: it is therefore difficult to measure the its centroid correctly. In addition, the line is at the opposite part of the spectrum relative to the interesting data and therefore our conclusion would be quite sensitive to a small error in the dispersion coefficients. We also compared the location of the Lyman-break visible in the individual spectra (see below). Unfortunately, the spectrum of the first orbit has half the exposure time of the subsequent ones. In addition, the spectra of the first and second orbits show structures probably caused by incorrect flat–fielding. We therefore found that the most reliable method to fix the wavelength calibration relative to the one of the first orbit is to use the location of the build-up at the red end of the spectra, caused by the decreasing dispersion of the PRISM (D. Lindler, private communication). The precision of the zero point calibration for the individual spectra is estimated to be better than 0.3 pixels. The reduced spectrum is shown in Fig. 1. A break is seen in the spectrum at $`\lambda 3000`$ Å. However, an excess of emission is seen shortward of it. We will come back to this point in §3.2. The first two rows of Table 2 gives the countrate and the integrated flux over 3000 Å $`<\lambda <`$ 3300 Å, as well as their errors, for each Prism spectrum, and the last row gives the mean counts. These values are uncorrected for Milky Way extinction. There is no evidence that the afterglow flux varied during the course of these observations (about 5 hours). The absence of flux recovery shortward of the Lyman edge at $`z=2.323`$ indicates a large H I column density. We return to this point in Sect. 2.1.3 below. The spectra are also affected by the HST and STIS Point Spread Function. Therefore we deconvolved the spectra using the method described in Smette et al. (2001). The integrated fluxes in the range 3000 Å $`<\lambda <`$ 3300 Å from the deconvolved spectrum along with the estimated errors are given in the last column of Table 2. #### 2.1.2 The G430L Spectra The G430L spectra cover the spectral range from 3000 Å to 5700 Å. The purpose of the observation was to determine the optical slope of the afterglow spectrum with good accuracy using near simultaneous observations over 3000–16000 Å together with the NICMOS data. As the GRB redshift was larger than foreseen when the Phase 2 was submitted, the spectra are affected by Ly$`\alpha `$ absorption (the Ly$`\alpha `$ forest starts at 4048 Å) and hence less useful for obtaining a precise measurement of the spectral slope. In addition, the spectra have a rather poor S/N and the telescope was not dithered between the 2 orbits as it was feared that the afterglow could be too faint for the processing of individual spectra. However, in order to limit the effect of the numerous cosmic rays, 2 CR–SPLIT exposures were made in each of the 2 orbits (cf. Table 1). Due to the lack of dithering, correct processing of hot and warm pixels is crucial but in practice very difficult. The reduced individual spectra still show spikes. They are usually one or two pixel wide and often common between the different spectra. A number of criteria were defined to decide that a given pixel in the combined spectrum is not valid (i.e., due to wrongly corrected warm or hot pixels) so that its value is not used in subsequent analysis: (a) its quality value is larger or equal to 175, which include among other conditions unrepairable hot pixels, pixels considered as cosmics by CR-SPLIT and saturated pixels, and hot pixels (CCD dark rate $`>`$ 0.2 counts/s); (b) pixels with ’net’ countrate larger than 0.09; and finally, (c) pixels with ’net’ countrate smaller than $`0.03`$ (blemishes). Figure 1 shows the resulting spectrum, its 1$`\sigma `$ error spectrum, as well as boxes representing the mean values of valid pixels within a 50 Å range, and corresponding standard error. These wavelength ranges were selected away from the absorption lines reported by Matheson et al. (2003). Their limiting wavelengths as well as their corresponding mean fluxes and errors, corrected from Milky Way extinction, are reported in Table 3. The best fitted power-law over these ranges has a slope of $`\beta =1.71\pm 0.14`$ with a reduced $`\chi ^2=0.73`$, significantly steeper than the $`\beta 1.0`$ in the X-ray band (see below). In Sect. 2.2.3 we analyze the full Spectral Energy Distribution (SED) and conclude that the steep slope in the UV most likely is due to (modest) Small Magellanic Cloud (SMC) like extinction in the host galaxy. It is worth noting that the integrated flux of the deconvolved Prism spectrum over 3000 Å $`<\lambda <`$ 3300 Å, corrected for Milky Way extinction is $`3.45\pm 0.11\times 10^{15}\mathrm{erg}\mathrm{cm}^2\mathrm{s}^1`$, while the corresponding value in the G430L spectrum is $`3.15\pm 0.46\times 10^{15}\mathrm{erg}\mathrm{cm}^2\mathrm{s}^1`$. The fact that these two values are consistent with each other gives us confidence that the flux calibration of both spectra is correct. #### 2.1.3 Limit on the H I Column Density Assuming that the underlying continuum can be represented by the extrapolation of the power-law described in the previous paragraph, we measure a total equivalent width of $`1.9\pm 0.5`$ Å using the 6 pixels covering the range from 2417 to 2560 Å (redwards of the Lyman-limit of the foreground $`z=1.60`$ Mg II absorber). The error is only the statistical error and it does not include the (likely larger) systematic error from the sky subtraction and flat-fielding. For a H I column density below $`1\times 10^{18}`$ cm<sup>-2</sup> at $`z=2.33`$, the continuum should have recovered, giving an equivalent width above 4.5 Å. Assuming no additional intervening strong (Lyman-limit) absorption, which is reasonable as there are no intervening metal line systems at $`1.60<z<2.33`$ (Møller et al. 2002), we conclude that the H I column density is larger than $`1\times 10^{18}`$ cm<sup>-2</sup> (5$`\sigma `$). The G430L spectra also show the Ly$`\alpha `$ absorption lines reported by (Møller et al. 2002), as well as the associated Ly$`\beta `$ lines. However, the latter appear in a noisy part of the spectrum, with a S/N/pixel of about 2.5, which is also affected by a number of bad pixels. Møller et al. (2002) find from the Ly$`\alpha `$ absorption line that the H I column density is constrained to be below $`1.1\times 10^{20}`$ cm<sup>-2</sup>. When we use the Voigt-profile parameters used to fit their low resolution spectrum (a total column density of $`6\times 10^{19}`$ cm<sup>-2</sup> and a $`b`$ parameter of 15 km s<sup>-1</sup>) we find that the corresponding Ly$`\beta `$ lines would appear weaker than the observed lines in the STIS spectrum. The significance of this finding is low due to the low S/N of the spectrum. However, we can conclude that, either the total column density is larger than $`6\times 10^{19}`$ cm<sup>-2</sup> or the $`b`$ parameter is larger 15 km s<sup>-1</sup>. Fiore et al. (2004) find, based on a high resolution spectrum obtained with the UVES spectrograph on the VLT, that the absorption profiles all contain very wide components showing that the latter explanation appears to be the correct one. From the UVES spectrum the H I column density is constrained to be in the range 1–10$`\times 10^{19}`$ cm<sup>-2</sup> (Castro-Tirado et al. 2005, in preparation). ### 2.2 ACS and NICMOS Imaging Observations The ACS observations from 2002 October 7 and 11 were CR-SPLIT observations consisting of only one or two exposures. These were reduced through the pipeline. The later epochs were observed with multiple exposures using non-integral multiple pixel dither steps and combined using MULTIDRIZZLE using pixfrac=1 and scale=0.66 (Fruchter & Hook 2002; Koekemoer 2002). All NICMOS images were taken using the NIC3 camera (51$`\times `$51 arcsec<sup>2</sup> field-of-view, 0.203 arcsec per pixel), MULTIACCUM mode (sample up the ramp), SPARS64 sample sequence and $``$3 arcsec dithers along each axis between exposures for bad pixel rejection. All data were taken in South Atlantic Anomaly free orbits except for the first two exposures of epoch 2. These were taken in an orbit following the last shallow pass of the day and were only mildly impacted, so correction for cosmic-ray persistence was not necessary. Two different calibration techniques were applied to this data. Epochs one, two and three (2002 October 7, 11 and 22, respectively) were calibrated with the standard NICMOS calibration pipeline CALNICA, using the best available reference files in the calibration database. Following CALNICA, the STSDAS task PEDSKY was used to remove any remaining quad-based DC biases (”pedestal”) signatures in the data. A sky image was constructed for each of the two filters using a source-masked median of all the epoch one and epoch two images combined, which were taken at different spacecraft orientations and dither positions. The sky image was then subtracted from each of the calibrated, PEDSKY corrected images. For epochs four and five (2003 May 26 and July 21) a different technique was used, employing new methods developed during the Hubble Ultra Deep Field (HUDF) NICMOS analysis. First, a correction was made to the raw images to remove amplifier crosstalk from bright sources (also known as the ”Mr. Stay-puft” anomaly). Although there are no bright sources in this field, every source produces a faint amplifier stripe and reflection to some level in the data. This is especially true of cosmic ray hits during the sample up the ramp MULTIACCUM exposures. Even though the cosmic ray hits themselves are flagged and rejected by the CALNICA processing, each hit leaves a faint stripe and reflection due to the amplifier effects, which can impact the noise floor. After this correction, the PEDSKY DC bias removal tool was run on dark-subtracted versions of each read up the ramp, and any measured bias offset (per quadrant) and sky background was subtracted from the raw reads. An image of the sky background from the HUDF was used as the sky model in pedsky rather than the usual internal lamp flat. The entire MULTIACCUM sequence was then run through CALNICA as usual, except that the HUDF superdark was used instead of the standard dark reference file. This was possible because the same readout sample sequence that was used for the HUDF (SPARS64) was also used for these data. #### 2.2.1 Photometry Photometry was done using aperture photometry. In Table 1 we provide measured countrates and the apertures sizes. For the 2002 October 7 and 11 observations the countrates in Table 1 are corrected for geometrical distortion using the Pixel Area Map from the STScI web-pages<sup>2</sup><sup>2</sup>2http://www.stsci.edu/hst/acs/analysis/zeropoints/analysis/PAMS. For the subsequent epochs this correction is done automatically during drizzling. To convert the measured countrates to magnitudes we used the synphot package under STSDAS. Synphot calculates the AB magnitude corresponding to the measured countrate and the applied aperture size. AB magnitudes derived in this way are given in the last column of Table 1. To accurately measure the afterglow component in the 2002 October 22 and 2002 November 26 F606W observations, where the host galaxy is contributing a large fraction of the flux, we drizzled these images onto the coordinate system of the 2003 May 31 image. By subtracting the May 31 image we could then measure the afterglow component using a small aperture with radius 0.2 arcsec<sup>3</sup><sup>3</sup>3We note that another transient source, presumably a supernova, is detected in the May 31 image superimposed on a faint galaxy at the celestial position RA(2000) = 00:26:55.8, Dec(2000) = 18:55:28.. For the NICMOS images we use a circular aperture with radius equal to 5 pixels for the early points (October) and 2.5 pixels for the late points (November and later) due to the larger pixels of the NIC3 detector (the pixel scale is 0.203 arcsec per pixel). To derive AB magnitude we multiply the counts with 1.075 (early) and 1.15 (late) to compensate for the finite aperture and use the most recent photometric keywords available at the HST web-site<sup>4</sup><sup>4</sup>4http://www.stsci.edu/hst/nicmos/performance/photometry/nic13\_postncs\_keywords.html to get the zero-points. #### 2.2.2 The Late-time Lightcurve The ACS photometry allows a precise measurement of the late-time decay slope of the optical/near-IR afterglow lightcurve. There is a substantial disagreement between different reports in the literature. The values of the late-time decay slope ranges from $`\alpha _2=1.43\pm 0.03`$ (H03 – note, however, that they argue for $`\alpha _2=1.98`$ as the best value when including also the broadband SED in the analysis) to $`\alpha _2=2.9`$ (M03). H03 find a jet-break time of $`4.74\pm 0.14`$ days, M03 find 9 days (no error bar) and Björnsson et al. (2004) 0.6 days. The main reasons for these large disagreements are the very “bumpy” lightcurve (making the determination of the epoch of the jet-break very sensitive to the sampling of the lightcurve) and the relatively bright underlying host galaxy (potentially complicating the determination of the late time slope $`\alpha _2`$). M03 assume an overly bright host galaxy magnitude (based on the November epoch of ACS observation in Table 1) and hence they derive an incorrectly large value of $`\alpha _2`$. Rol et al. (2003) argue for a jet-break around 1 day based on the polarimetry, whereas Lazzati et al. (2004) find that a later jet-break time of about 3$`\pm `$1 days is consistent with the polarimetry after day 1. Björnsson et al. (2004) are able to fit both the optical lightcurve, the radio and X-ray observations and the full evolution of the polarization using a model with four episodes of energy injection. In this model the real jet break takes place around 0.6 days and the apparent later break time is an artifact of the rebrightenings caused by energy injections. The epoch 5 observation from May 31 2003 is more than half a year after the explosion, and the afterglow light at this epoch is most likely negligible. Conservatively assuming a late-time decay-slope of $`\alpha _2=1.7`$ we infer an afterglow magnitude of 28.5 in the F606W band at 2003 May 31. The decay-slope $`\alpha _2`$ is larger than 1.7 (see below) and the afterglow magnitude at 2003 May 31 will therefore be fainter than 28.5. The rightmost column in Table 1 contains the afterglow magnitudes based on the assumption of no afterglow emission in the last epoch images. In Fig. 3 we plot the F606W magnitudes for epochs 1–4. Clearly, the four points are not consistent with a single power-law. Epoch 1 is before the break time as determined both by H03 and M03 and we can hence exclude this point from the power-law fit. However, a single power-law fit to the last three points is also formally rejected ($`\alpha _2=1.86`$, $`\chi ^2=9`$, 1 d.o.f). The dotted line shows the result of an unweighted fit to the epoch 2–4 points, and the dashed line shows the line defined by the epoch 2 and 3 points. The dotted and dashed lines correspond to $`\alpha _2=1.77`$ and $`\alpha _2=1.88`$, respectively. The last HST detection indicates a flattening of the late-time afterglow. This effect cannot be caused by residual afterglow emission in the epoch 5 image as this would have the opposite effect, making the lightcurve bend the other way. The flattening could be the result of the early transition to non-relativistic expansion. According to Livio & Waxman (2000) $`\alpha _2`$ should evolve towards $``$0.9 on a timescale of 5 months. An underlying supernova (Stanek et al. 2003; Hjorth et al. 2003a) is not expected to contribute significantly to the restframe UV light corresponding to the ACS filters. An alternative explanation is a bump in the light curve around 2002 October 11 causing the afterglow to be too bright. Given the very bumpy nature of the GRB 021004 lightcurve this does not appear unlikely. #### 2.2.3 The NIR/Optical SED of the Afterglow As shown in Fig. 5, the HST observations allow us to construct the SED of the afterglow around 2002 October $``$7.20 and $``$11.67. Given that the afterglow photometry is based on images where the host galaxy has been subtracted, the SED is not affected by host galaxy light contamination. The HST photometry was shifted to a common epoch at 2002 October 7.20 and 11.67 assuming power-law decay slopes of $`\alpha =0.85`$ (from H03) and $`\alpha =1.86`$ (based on the analysis described above), respectively. The two observation sets are well clustered around these two common dates (maximum epoch shifts of $`\delta t=`$ 0.28 days for October $``$7.20), so the derived SEDs are not very sensitive to the assumed values of $`\alpha `$ before the lightcurve break at $``$7.20 (e.g., $`\delta m<0.16`$ for $`0.5<\alpha <2.0`$). We have added in quadrature a 0.05 mag error to the F110W-band magnitudes displayed in Table 1 in order to account for the NIC3 intra-pixel sensitivity variations. These sensitivity variations are especially relevant in the F110W-band and can produce sensitivity variations as large as $`30\%`$ (peak-to-peak, Storrs et al. 1999). We analyze the SED by fitting a function of the form $`F_\nu \nu ^\beta \times 10^{0.4A_\nu }`$, where $`\beta `$ is the spectral index and $`A_\nu `$ is the extinction at frequency $`\nu `$. $`A_\nu `$ has been parametrized in terms of $`A_\mathrm{V}`$ following the three extinction laws Milky Way (MW), Large Magellanic Cloud (LMC), and SMC given by Pei et al. (1992). In the analysis we have included the UV flux from the STIS G430L spectrum (Table 3). The Pei et al. (1992) extinction laws are very uncertain for restframe wavelengths below 1000 Å, so the bluest bins of the STIS G430L spectrum were not included in the analysis. For the same reason, and also due to the Ly$`\alpha `$ blanketing and Lyman-limit absorption, the F250W data point was also not included in the SED fit. As a consistency check we have corrected the F250W data point for Ly$`\alpha `$ blanketing by convolving the F250W filter sensitivity curve with a Ly$`\alpha `$ blanketing model and a Lyman-limit break corresponding to the burst redshift. The two excluded data points are roughly compatible with the optical/NIR SED (the two stars in the upper panel of Fig. 5). For both epochs the SEDs are clearly curved and hence highly inconsistent with a pure power-law spectrum ($`\chi ^2/\mathrm{d}.\mathrm{o}.\mathrm{f}.16.7`$, see Table 4). The LMC, and especially the MW extinction law which yields $`A_\mathrm{V}<0`$, provide unacceptable fits. Only the SMC extinction law provides moderate $`\chi ^2/\mathrm{d}.\mathrm{o}.\mathrm{f}.`$ values. The $`A_\mathrm{V}`$ and $`\beta `$ values derived for the SMC ($`\beta =0.30\pm 0.06`$, $`A_\mathrm{V}=0.23\pm 0.02`$ at October 7.20 and $`\beta =0.42\pm 0.06`$, $`A_\mathrm{V}=0.20\pm 0.02`$ at October 11.67) are consistent with the ones reported by H03 at 2002 October 10.072: $`\beta =0.39\pm 0.12`$, $`A_\mathrm{V}=0.26\pm 0.04`$. The evolution of the afterglow is achromatic within the errors from 4000 Å to 16000 Å over the period from October 7 to October 22 (3 to 18 days after the burst). The F606W$``$F160W color from November 25–26 (53 days after the burst) observation is consistent with the color from the earlier epochs, but the error-bar is very large. H03 find that R$``$I is constant from 0.35 days to 5.5 days after the burst. On the other hand, Matheson et al. (2003) and Bersier et al. (2003) present evidence for color changes in the near-UV and optical range during the first four nights. The only way to reconcile these observations with ours is that the color changes must be related to a stochastic phenomenon superimposed on the normal afterglow light whose color remains very stable. Moreover, the cooling break cannot have passed through the optical range during the period from 0.35 to 18 days after the burst and possibly not earlier than 53 days. #### 2.2.4 Broadband SED The X-ray afterglow of GRB 021004 was observed with the Chandra X-ray Observatory on 2002 October 5.9 and 2002 November 25 (Sako & Harrison 2002a,b) and is analyzed in H03 and Fox et al. (2003). The X-ray spectra show no evidence for absorption in addition to the foreground Galactic absorption with a 90% confidence upper limit on the absorbing column density in the host galaxy or along the line-of-sight of 2.3$`\times `$10<sup>21</sup> cm<sup>-2</sup>. H03 report a spectral slope in the X-rays of $`\beta _X=1.06\pm 0.06`$ and $`\beta _X=0.94\pm 0.03`$ fitting to 2–10 keV and 0.4–10 keV respectively. On November 25 (52.3 days after the burst) the X-ray flux was reduced to $`7.2\pm 2.5\times 10^{16}`$ erg cm<sup>-2</sup> s<sup>-1</sup> compared to $`4.3\times 10^{13}`$ erg cm<sup>-2</sup> s<sup>-1</sup> at around 1.5 days (Fox et al. 2003). This corresponds to a temporal decay slope in the X-rays of $`\alpha _X=1.80\pm 0.10`$ – very similar to what we find for the late-time slope in the optical/near-IR. This means that the broad-band SED from the optical to the X-ray band has a roughly constant shape $``$2 days and $``$52 days after the burst. In Sect. 3 we return to the interpretation of the broad-band SED within the context of the blast wave (fireball) and cannonball models. In the lower panel of Fig. 5 the broadband SED from radio to X-rays is plotted. The epochs cluster around 2002 October 10.7, 11.67, and 2002 November 25.73. All the radio measurements and the X-ray flux of November 25.73 are taken from the literature. The X-ray points of October 7.20 (empty squares in the lower panel of Fig. 5) are based on our independent analysis of the Chandra X-ray spectrum taken on Oct 5.4–6.4 and shifted in time to October 7.20 assuming a decay of $`\alpha =1.86`$ as derived from the optical light curve. #### 2.2.5 The Host Galaxy The host galaxy is clearly detected in all bands in the latest epoch images from 2003 May–July. In the left panel of Fig. 5 we show a 1$`\times `$1 arcsec<sup>2</sup> section of the F606W image from 2003 May 31 around the host galaxy. The host galaxy has a very compact core with a half light radius of only 0.12 arcsec. The resolution of the ACS image is about 0.05 arcsec so the core is well resolved. The galaxy has a faint second component offset by about 0.28 arcsec (2.2 kpc) towards the east. We cannot, based on the imaging alone, determine if this second component is part of the host or due to an independent object, e.g. one of the foreground absorbers. The position of the afterglow is marked with a cross and the 5$`\sigma `$ error is shown with a circle (1$`\sigma `$ = 0.08 pixels). The afterglow position is offset from the galaxy core by only 0.4 drizzled pixels, corresponding to 0.015 arcsec or 119 parsec at $`z=2.33`$. This is one of the smallest impact parameters measured so far (Bloom, Kulkarni & Djorgovski 2002), and although it could be a chance projection it suggests that the progenitor could be associated with a circumnuclear starburst. The right panel of Fig. 5 shows a larger portion, 10$`\times `$10 arcsec<sup>2</sup>, of the field around the host galaxy. A number of very faint (R $`>`$ 26) galaxies are seen. Some of these are likely associated with the strong intervening Mg II absorbers at redshifts $`z=1.38`$ and $`z=1.60`$ seen in the afterglow spectrum (Møller et al. 2002; Castro-Tirado et al. 2004; see also Vreeswijk, Møller & Fynbo 2003; Jakobsson et al. 2004). The SED of the galaxy can be quite well constrained from restframe 1200 Å to 5000 Å based on the ACS and NICMOS detections. This range brackets the UV region sensitive to young O and B stars, and hence the star formation rate, and the Balmer jump, sensitive to the age of the starburst. In Fig. 6 we show the SED of the host. The SED is essentially flat bluewards of the Balmer jump, and the Balmer jump has a magnitude of about 0.4 mag. Also shown is the simulated spectrum of a starburst from the 2003 version of the spectral synthesis models of Bruzual & Charlot (2003). The model starburst has an age of 42 Myr and a metallicity of 20% solar. It is possible to obtain reasonable fits with ages between 30 and 100 Myr dependent on the metallicity. In addition the range of allowed models could be increased by allowing for extinction, but as the galaxy is known to have high equivalent width Ly$`\alpha `$ line emission (Møller et al. 2002) there is little room for extinction (e.g., Charlot & Fall 1993). Based on the strength of the UV continuum of the host galaxy we can derive a value of the Ly$`\alpha `$ equivalent width (Møller et al. 2002 give a lower limit). The magnitude of the host galaxy of F606W<sub>AB</sub> = $`24.39\pm 0.04`$ corresponds to a specific flux of $`F_\lambda =1.1\times 10^{18}`$ erg s<sup>-1</sup> cm<sup>-2</sup> Å<sup>-1</sup>. With the measured line-flux of $`F_{\mathrm{Ly}\alpha }=2.5\pm 0.5\times 10^{16}`$ erg s<sup>-1</sup> cm<sup>-2</sup> we derive an observed equivalent width of $`231\pm 47`$ Å corresponding to $`69\pm 14`$ Å in the restframe, very similar to what is found for the host galaxy of GRB 000926 at $`z=2.04`$ (Fynbo et al. 2002). #### 2.2.6 Star Formation Rate The specific luminosity of the UV continuum provide a measurement of the SFR. We use the relation of Kennicutt (1998) $$SFR(\mathrm{M}_{\mathrm{}}\mathrm{yr}^1)=1.4\times 10^{28}L_\nu ,$$ where $`L_\nu `$ is the specific luminosity in units (erg s<sup>-1</sup> Hz<sup>-1</sup>) in the 1500–2800 Å range. The observed AB magnitude of 24.55 corresponds to $`F_\nu =5.50\times 10^{30}\mathrm{erg}\mathrm{s}^1\mathrm{Hz}^1\mathrm{cm}^2`$. In our choice of cosmology this corresponds to a value of $`L_\nu =4\pi d_l^2F_\nu /(1+z)=6.93\times 10^{28}\mathrm{erg}\mathrm{s}^1\mathrm{Hz}^1`$. This gives a SFR of $`10\mathrm{M}_{\mathrm{}}\mathrm{yr}^1`$. Due to dust absorption this is a lower limit to the SFR, but as mentioned above there is only little room for dust. We can also derive an estimate for the SFR based on the Ly$`\alpha `$ flux. Møller et al. (2002) report a Ly$`\alpha `$ flux of $`2.46\pm 0.50\times 10^{16}\mathrm{erg}\mathrm{s}^1\mathrm{cm}^2`$. This corresponds to a luminosity of $`1.0\pm 0.2\times 10^{43}\mathrm{erg}\mathrm{s}^1`$ and a SFR of $`10\mathrm{M}_{\mathrm{}}\mathrm{yr}^1`$ (following the calculation in Fynbo et al. 2002). A similar value was found by Djorgovski et al. (2002). The values of the SFR measured from the UV continuum and from the Ly$`\alpha `$ flux are hence in good agreement as for GRB 030323 (Vreeswijk et al. 2004). Hence, at least for some high-$`z`$ starbursts Ly$`\alpha `$ emission is as reliable as the UV-continuum as SFR estimator (see Mas-Hesse et al. 2003 and Kunth et al. 2003 for critical discussions of the use of Ly$`\alpha `$ as SFR indicator). ## 3 Discussion and Conclusions ### 3.1 The H I Column Density The STIS spectroscopy is consistent with and complements a the measurements from ground based near-UV and optical spectroscopy. We detect the Lyman-limit break associated with the GRB 021004 absorption system also detected in the Ly$`\alpha `$ resonance line and several metal lines at observer-frame optical wavelengths. Based on the Lyman-limit break we have placed a lower limit of about $`1\times 10^{18}`$ cm<sup>-2</sup>, significantly higher than the $`2\times `$10<sup>16</sup> cm<sup>-2</sup> inferred by M03 based on the Lyman series. The low column densities derived by M03 are caused by their assumption that the lines are not saturated. From the analysis of the X-ray afterglow Fox et al. (2003) derive an 90% confidence upper limit on the absorbing column density in the host galaxy or along the line-of-sight of 2.32$`\times `$10<sup>21</sup> cm<sup>-2</sup>. The H I column density is quite low compared to that found for other GRB host galaxies (Vreeswijk et al. 2003, their Fig. 4 and discussion thereof). Most of the Ly$`\alpha `$ lines from GRB afterglows are strongly damped and have inferred column densities well above 10<sup>21</sup> cm<sup>-2</sup>. Only GRB 021004, GRB 011211 (Vreeswijk et al. 2005) and possibly GRB 030226 (Klose et al. 2004) depart from this rule having column densities below the classical limit of 2$`\times `$10<sup>20</sup> cm<sup>-2</sup> H I for Damped Ly$`\alpha `$ Absorbers (DLAs) in QSO spectra. The Ly$`\alpha `$ absorption profile is furthermore peculiar by having multiple components spread over about 3000 km s<sup>-1</sup> with clear evidence for line-locking between the components (Savaglio et al. 2002; Møller et al. 2002; Schaefer et al. 2003; M03; Castro-Tirado et al., in preparation). The line-locking implies a very strong radiation field along the line of sight, presumably from the GRB progenitor star/star cluster, and it is possible that this radiation field has ionized the H I along the line thereby explaining the low total column density in the host galaxy. An alternative explanation for the low foreground column density compared to previously studied GRB lines-of-sight could be that the progenitor was located in the perimeter of the host galaxy. However, the position of the afterglow very close to the host (although in projection) in Fig. 5 does not support this explanation. ### 3.2 A Ly-$`\alpha `$ emission line at $`z=1.38`$? We noted in Sect. 2.1.1 that the PRISM spectrum seems to show an excess of emission shortward of the Lyman-break. As the different methods to derive the HI column density at the GRB redshift are consistent with each other, it is unlikely that we overestimate its value. Therefore, a low HI column density cannot explain the shape of the Lyman-break. One possibility is that this excess emission could be caused by a Ly$`\alpha `$ emission line at $`z=1.38`$, probably associated with absorption line system A following the naming convention in Møller et al. (2002). In order to examine if this explanation is plausible, we have subtracted a model of the afterglow spectrum based on the results in Sect. 2.2.3 from the observed spectra. In addition, H I continuous absorption has been added to model the Lyman-break, with an H I column density $`\mathrm{log}N_{\mathrm{HI}}>18`$. This model is then used as an input to the SIM\_STIS IDL routine written by P. Plait (private communication), which accurately models the behavior of STIS. The output of this routine is then convolved by the expected PSF and subtracted from the observed spectra. The residuals for the four individual spectra are shown in Fig. 2. The S/N of the emission line over the mean of the 4 individual spectra is 23. This figure also shows for comparison the spectrum of a $`6\times 10^{16}\text{ergs}\text{s}^1\text{cm}^2`$, $`\lambda =2890\text{Å}`$ unresolved emission line convolved by the HST+STIS LSF. Such a line can indeed be identified as Ly$`\alpha `$ at $`z=1.38`$. If due to Ly$`\alpha `$ at $`z=1.38`$ this flux would correspond to a Star Formation Rate (SFR) of about 7 $`\mathrm{M}_{\mathrm{}}\mathrm{yr}^1`$ following the calculation in Fynbo et al. (2002). This is formally inconsistent with the 2$`\sigma `$ upper limit on the SFR of 2.3 $`\mathrm{M}_{\mathrm{}}\mathrm{yr}^1`$ based on the O III line in Vreeswijk, Møller & Fynbo (2003), but the systematic uncertainties in the relations between Ly$`\alpha `$ and O III luminosities and SFR are significant enough to leave some room for this interpretation, eventhough we consider it unlikely. If the line is due to the $`z=1.38`$ absorber its impact parameter relative to the GRB line-of-sight would have to be very small, a tenth of an arcsec at most and consistent with zero. Similar small impact parameters have been observed for DLAs (e.g., Møller, Fynbo & Fall 2004). Savaglio et al. (2002) note that the absorbtion properties of the $`z=1.38`$ absorber suggests that it is a DLA. ### 3.3 Reddening Despite the relatively low H I column density we detect statistically significant reddening in the afterglow SED. This is surprising if the dust properties of the ISM are similar to those of SMC dust like for previously studied GRB afterglows (e.g., Hjorth et al. 2003b). For the SMC extinction curve, for which R$`{}_{\mathrm{V}}{}^{}=2.93`$ (Pei 1992), the observed extinction corresponds to E(B$``$V)=0.07. For a H I column density of 1–10$`\times 10^{19}`$ cm<sup>-2</sup> we derive a gas-to-dust ratio of N(H I)/E(B$``$V) = 1.5–15$`\times 10^{20}`$ cm<sup>2</sup> mag<sup>-1</sup>. This is significantly more reddening per column density than for SMC dust, where N(H I)/E(B$``$V) = 4.4$`\pm `$0.7$`\times 10^{22}`$ cm<sup>2</sup> mag<sup>-1</sup> (Bouchet et al. 1985). This could imply that a large fraction of the hydrogen along the line of sight is ionized. Alternatively, some of the reddening could be due to dust in the two foreground Mg II absorbers, but this appears unlikely as foreground absorbers in general cause very little reddening (Murphy & Liske 2004). In other words, the column densities of the foreground absorbers would have to have higher column densities than any known DLA, and this is very unlikely. The intrinsic shape of the afterglow SED is not expected to be a pure power-law as the cooling break is located very close to the near-UV/optical bands (see below). This may explain some of the difference, but we do not expect all the observed bending of the SED to be due to the intrinsic shape of the SED (see, e.g., Granot & Sari 2002). ### 3.4 Comparison with Afterglow Models The post-break decay and spectral slope have been well constrained with the data presented in this paper. The spectral slope $`\beta `$ in the optical to near-IR range corrected for extinction is $`0.36`$ (mean value of the two measurements in Table 4). The late-time decay slope $`\alpha _2`$ is in the range $`\alpha _2=1.8`$–1.9. These parameters agree well with the results found by H03. From these afterglow parameters we can infer the position of the cooling frequency $`\nu _c`$ relative to optical frequencies $`\nu _\mathrm{O}`$ in the context of the standard blastwave model (see Mészáros 2002 for a review). For $`\nu _c>\nu _\mathrm{O}`$ we expect $`\alpha _22\beta =1`$, and for $`\nu _c<\nu _\mathrm{O}`$ we expect $`\alpha _22\beta =0`$ (Sari et al. 1999; Chevalier & Li 1999). Our observations imply $`\alpha _22\beta =1.0`$–1.1, clearly indicating that $`\nu _c>\nu _\mathrm{O}`$. This is consistent with the fact that the spectral slope in the X-rays, $`\beta _X=1.0`$, is significantly larger than in the optical, which makes a spectral break between the optical and X-ray bands unavoidable. The change in slope of $`\mathrm{\Delta }\beta 0.5`$ is exactly what is expected across the cooling break, $`\mathrm{\Delta }\beta =1/2`$. We note that the strong bumps in the optical (and possibly X-ray lightcurves, Fox et al. 2003) can only be interpreted as being the result of density fluctuations if the cooling break is bluewards of the optical (and X-rays) (Lazzati et al. 2002, but see also Nakar et al. 2003; Nakar & Piran 2003; Björnsson et al. 2004). There is one complication with this conclusion: the number $`N(E)`$ as function of the electron energy $`E=\gamma _em_ec^2`$ of the electrons producing the synchrotron emission is expected to have the form $`N(E)E^p`$, and for $`\nu _c>\nu _\mathrm{O}`$ the theory predicts $`\alpha _2=p`$. This means that $`p1.9`$, which results in a divergent energy spectrum. This case has been analyzed by Dai & Cheng (2001) and Bhattacharya (2001), but as $`p`$ is in our case very close to 2 the resulting relations between decay and spectral slopes are quite similar to the equations for the $`p>2`$ case. The fact that the optical near-IR colors of the afterglow remain constant means that the cooling break has to be located on the blue side of the optical during the period from 0.35 days after the burst and at least until October 22 and possibly longer than November 25–26. Depending on the geometry of the blastwave and the density profile of the surrounding medium the cooling break will move towards higher frequencies (wind environment, Chevalier & Li 1999), lower frequencies (ISM, spherical geometry) or remain constant (ISM, jet geometry, Sari et al. 1999; Chevalier & Li 1999). The fact that the optical to X-ray slope $`\beta _{\mathrm{OX}}`$ has remained constant within the errors from October 7 and until November 26 with a value close to $`\beta _X`$ means that the cooling frequency must have remained constant and close to the optical in this time span. This implies a jet geometry and a constant density environment. Li & Chevalier (2003) argue that an apparent break in the very early lightcurve ($`t<0.1\mathrm{days}`$) is best understood assuming a wind-shaped circumburst medium. The only way to reconcile this with the discussion above is that the apparent break in the optical lightcurve is due to lightcurve fluctuations rather than being due to the typical frequency $`\nu _m`$ passing down from higher frequencies as suggested in the wind model of Li & Chevalier (1999). The cannonball model offers an alternative explanation for the GRB phenomenon (Dado, Dar & De Rújula 2002). In this model the jet opening angle is much smaller, and the relativistic $`\gamma `$ factor higher than in the fireball model. Dado, Dar & De Rújula (2003) have presented an analysis of GRB 021004 in the cannonball model in which a very good fit to most of the available groundbased data up to about 30 days after the burst was obtained. In the cannonball model the asymptotic (late-time) behavior of the afterglow is $`F_\nu (t)t^{2.13}\nu ^{1.1}`$ (Dado et al. 2002). The effective decay slope for the cannonball model fit GRB 021004 in Dado et al. (2003) between October 11 and November 25 is $`\alpha _2=1.92`$, which is close to the observed value. A spectral slope change towards $`\beta =1.1`$ is not observed up to 50 days after the burst. ### 3.5 The Host Galaxy The host galaxy is extremely blue in the observed optical bands. Fig. 7 shows the optical colors of the host in a color-color diagram. The points with error bars represent galaxies in the environment of the host (from 70$`\times `$70 arcsec<sup>2</sup> around the host) and the dots are colors of galaxies from the ACS observations of the GOODS South field<sup>6</sup><sup>6</sup>6For the GOODS data we use the average of F775W and F850LP magnitudes as a proxy for F814W. As seen, the host is in the extreme blue color of the distribution for field galaxies. The reason for this is a combination of a young age for the star burst ($`<`$100 Myr) and the redshift causing a very small Ly$`\alpha `$ blanketing in the F435W filter and placing the Balmer jump beyond the F814W band. The optical bands hence all probe the restframe UV continuum of the newly formed massive stars in the galaxy. A SFR of about 10 M yr<sup>-1</sup> is derived from the restframe UV flux density. The SFR inferred from the Ly$`\alpha `$ luminosity is consistent with this value. The host galaxy has been observed in the sub-mm range with SCUBA (Tanvir et al. 2004), but it was not detected above a 2$`\sigma `$ limit of 2.5 mJy. All evidence is consistent with the host being a young, dust-poor starburst. This is typical for GRB host galaxies (Fruchter et al. 1999; Le Floc’h et al. 2003; Christensen et al. 2004; Courty, Björnsson & Gudmundsson 2004; Jakobsson et al. 2005). So far, 15 GRBs have been detected at redshifts where Ly$`\alpha `$ is observable from the ground (see Table 4 in Jakobsson et al. 2005). Of these, GRB 021004 has the intrinsically brightest detected host galaxy. Nevertheless, it is not brighter than the characteristic luminosity L for Lyman-break galaxies at slightly larger redshifts (e.g., Adelberger & Steidel 2000). The reason why most GRB host galaxies are relatively faint, dust-poor starbursts is not yet established. It could be that most of star-formation at these redshifts are located at the faint end of the luminosity function. There is evidence that the faint end slope of the luminosity function at high redshift is significantly steeper than in the local Universe (Adelberger & Steidel 2000; Shapley et al. 2001), so this is not unlikely (Jakobsson et al. 2005). However, there is still substantial uncertainty about the faint end slope (e.g., Gabasch et al. 2004). Another possibility is a low metallicity preference for GRBs as predicted by the collapsar model (MacFadyen & Woosley 1999; see also Fynbo et al. 2003; Le Floc’h et al. 2003, Prochaska et al. 2004). With the current very inhomogeneous sample of GRB host galaxies we cannot exclude that the current sample is somewhat biased against dusty starbursts (see also Ramirez-Ruiz, Trentham & Blain 2002). The currently operating Swift mission offers the possibility to resolve this issue (Gehrels et al. 2004). We thank Arnon Dar and Gunnlaugur Björnsson for critical comments on earlier versions of the manuscript and the anonymous referee for a thorough and constructive report. Support for Proposal number GO 9405 was provided by NASA through a grant from the Space Telescope Science Institute, which is operated by the Association of Universities for Research in Astronomy, Incorporated, under NASA contract NAS5-26555. We thank the schedulers of the Space Telescope, and in particular our Program Coordinator, Ray Lucas, for the extraordinary effort they put in to assure timely observations. AJL acknowledges support from the Space Telescope Science Institute Summer Student Programme and from PPARC, UK. STH acknowledges support from the NASA LTSA grant NAG5–9364. PJ acknowledges support from a special grant from the Icelandic Research Council. This work was conducted in part via collaboration within the the Research and Training Network “Gamma-Ray Bursts: An Enigma and a Tool”, funded by the European Union under contract number HPRN-CT-2002-00294. This work was also supported by the Danish Natural Science Research Council (SNF), the Carlsberg Foundation, and by the Danish National Research Foundation.
warning/0506/astro-ph0506168.html
ar5iv
text
# A Candidate Neutron Star Associated with Galactic Center Supernova Remnant Sagittarius A East ## 1 INTRODUCTION Sagittarius (Sgr) A East, a part of the Sgr A complex in the Galactic center, is an extended radio source with an angular size of 2$`.^{}`$7 $`\times `$ 3$`.^{}`$6 (Ekers et al., 1983). The nonthermal radio spectrum (radio spectral index $`\alpha `$ = 0.76, where S$``$$`\nu `$) and the shell-like morphology suggested an identification of a supernova remnant (SNR) for Sgr A East (Jones, 1974; Ekers et al., 1975, 1983). The ROSAT, ASCA, and Beppo-SAX satellites detected diffuse X-ray emission from Sgr A East, but extensive analysis was infeasible because of the limited detector capabilities (Predehl & Trümper, 1994; Koyama et al., 1996; Sidoli et al., 1999). Recently, observations with the high angular resolution instruments aboard the Chandra X-Ray Observatory clearly resolved the diffuse X-ray emission interior to the Sgr A East radio shell (Maeda et al. 2002, M02 hereafter). The Chandra data revealed a centrally-peaked X-ray morphology of Sgr A East. The X-ray emission originates from a hot thermal plasma with an electron temperature $`kT`$ $``$ 2 keV, showing strong He$`\alpha `$ lines from highly-ionized elemental species of S, Ar, Ca, and Fe. The estimated low mass of the hot gas ($``$2 M) and the total thermal energy ($`E_{th}`$ = 2 $`\times `$ 10<sup>49</sup> ergs) indicated a single SNR origin for Sgr A East (M02). The Chandra data revealed that the Fe-rich ejecta is concentrated near the center of the SNR, while other elemental species are uniformly distributed over the SNR. M02 further suggested that Sgr A East is a metal-rich, mixed-morphology (MM) SNR from a Type II SN explosion of a 13$``$20 M progenitor. Sgr A East was also observed with the XMM-Newton Observatory (Sakano et al. 2004, S04 hereafter). The results from the XMM-Newton data analysis were also consistent with a single SNR interpretation for Sgr A East. S04, however, found that a two-temperature thermal plasma ($`kT`$ $``$ 1 keV and 4 keV) was required in order to adequately describe the observed X-ray emission line features. The measured metal abundance patterns and the estimated Fe ejecta mass suggested either Type Ia or Type II origin. A weak 6.4 keV “neutral” Fe line feature was also detected with the XMM-Newton data. This neutral Fe line was attributed to the irradiation by Sgr A East of the molecular clouds which are also interacting with Sgr A East (S04). The discrepancies between the Chandra and the XMM-Newton data might have been caused by the low photon statistics of the Chandra data (S04). Effects of the relatively poor angular resolution of the XMM-Newton data may not be ruled out, either. Since the early Chandra observations of Sgr A East presented by M02, we have performed follow-up observations of the Galactic center with Chandra as parts of the monitoring program of the Galactic center supermassive black hole candidate Sgr A\* (Baganoff et al., 2003). As of 2002 June, the total exposure of Chandra observations of the Galactic center reached $``$600 ks, representing the deepest X-ray observations of the Galactic center. Utilizing the wealth of these deep observations, we have presented the complex nature of the pointlike and diffuse X-ray emission features in the Galactic center in a series of publications (Baganoff et al., 2003; Muno et al. 2003a, ; Muno et al. 2003c, ; Park et al. 2004a, ; Muno et al. 2004a, ; Muno et al. 2004b, ). The current deep Chandra data increase the photon statistics of Sgr A East by an order of magnitude, compared with that of M02. Taking advantage of the deep Chandra exposures, in this paper we present spatially-resolved spectral analysis of Sgr A East, which was infeasible with previous data. This approach is effective for the analysis of the ejecta material, excluding substantial contamination from the swept-up ISM, as successfully demonstrated with other Galactic SNR studies (e.g., Park et al. 2004b). The large-scale, global spectral analysis of Sgr A East has been performed with the previous Chandra and XMM-Newton observations, which revealed overall spectral parameters and thermal characteristics of the SNR (M02; S04). Rather than presenting such general characteristics, we concentrate in this work on the origin of Sgr A East in the context of the SN explosion types. Particularly, we draw an attention to a hard pointlike source detected in the northern edge of Sgr A East. The overall X-ray characteristics suggest that this source is a candidate, high-velocity neutron star (NS) associated with the SNR Sgr A East. The observations are briefly described in §2. The X-ray image and spectral analyses of the SNR Sgr A East and the candidate NS are presented in §3 and §4, respectively. We discuss key characteristics of the SNR and the NS candidate in §5. Finally, a summary is presented in §6. ## 2 OBSERVATIONS & DATA REDUCTION Since 1999, the Galactic supermassive black hole candidate Sgr A\* has been monitored with the Advanced CCD Imaging Spectrometer (ACIS) (Garmire et al., 2003) on board the Chandra X-Ray Observatory (Table 1). As of 2002 June, combining 11 Chandra/ACIS observations (except for ObsID 1561a, which was severely contaminated by a bright transient source within the field of view \[FOV\]; Muno et al. 2003b), the total exposure has reached $``$590 ks, which is the deepest ever observation of the Galactic center region in X-rays. We utilized the data reduced by Muno et al. (2003a), as we briefly describe here. We first applied the algorithm developed by Townsley et al. (2002a) to correct the spatial and spectral degradation of the ACIS data caused by radiation damage, known as charge transfer inefficiency (CTI; Townsley et al. 2000). We then screened the data by status, grade, and the flight timeline filter. We have also removed observation time intervals of strong flaring in the background. All individual event files were then reprojected to the tangent plane at the radio position of Sgr A\* (RA\[J2000\] = 17<sup>h</sup> 45<sup>m</sup> 40<sup>s</sup>.0409, Dec\[J2000\] = $``$29 00 28$`\stackrel{}{\mathrm{.}}`$118) in order to generate the composite data. The detailed descriptions of the data reduction process and the resulting broadband raw image from the composite data may be found in Muno et al. (2003a). ## 3 Sgr A East ### 3.1 X-Ray Images The center of Sgr A East radio shell is $``$50<sup>′′</sup> northeast of Sgr A\* and thus the entire SNR Sgr A East was imaged within the 17 $`\times `$ 17 FOV of the ACIS-I array during the Sgr A\* monitoring observations. A “true-color” X-ray image of the Sgr A East is presented in Figure 1. Each subband image has been exposure-corrected utilizing the exposure map produced by Muno et al. (2003a), and adaptively smoothed to achieve signal-to-noise (S/N) ratio of 4 by using the CIAO tool csmooth. Sgr A\* is marked near the center of the image. The bright X-ray feature around Sgr A\* is emission from a massive star cluster within the inner parsec of the Galactic center. The bright diffuse X-ray emission to the immediate east of Sgr A\* is the SNR Sgr A East. As reported with previous Chandra observations by M02, X-ray emission from Sgr A East is centrally enhanced with no apparent shell-like features (Figure 1). The enhancements of the blue emission around the center of the SNR are remarkable. As we discuss in the next section, this enhanced hard X-ray emission is primarily from the Fe He$`\alpha `$ line emission. The outskirts of the SNR show soft X-ray emission. Also evident is red, soft X-ray emission extending toward the northern side of the SNR, the so-called “plume” (M02; Baganoff et al. 2003). M02 suspected the existence of a high-velocity neutron star at the “tip” of the plume, which is physically associated with Sgr A East. With the deep exposure, a hard pointlike source CXOGC J174545.5$``$285829 was indeed detected there (Muno et al. 2003a, ). This hard pointlike source might thus be the predicted high-velocity neutron star candidate, which might have also produced the bow shock-like plume: hereafter, we name this NS candidate the “cannonball”. We discuss the cannonball in detail in § 4 and § 5.1. ### 3.2 Equivalent Width Images In order to examine the overall distributions of the diffuse emission line features from Sgr A East, we construct equivalent width (EW) images for the detected atomic emission lines, following the method described in Park et al. (2004a). After removing all detected point sources from the broadband image (see Muno et al. \[2003a\] for the details of the point source detection), subband images for the line and continuum bandpasses were extracted for each spectral line of interest. These subband images were adaptively smoothed to achieve an S/N ratio of 3. The underlying continuum was calculated by logarithmically interpolating between images made from the higher and lower energy “shoulder” of each broad line. The estimated continuum flux was integrated over the selected line width and subtracted from the line emission. The continuum-subtracted line intensity was then divided by the estimated continuum on a pixel-by-pixel basis to generate the EW images for each element. In order to avoid noise in the EW maps caused by poor photon statistics near the CCD chip boundaries, we have set the EW values to zero where the estimated continuum flux is low. As discussed in Park et al. (2004a), contaminations from the cosmic X-ray background and the particle background, both of which have not been subtracted in the EW generation, are insignificant in the EW images. Although the adaptive smoothing may introduce spurious faint features, the bright, arcminute-scale features are confirmed by our spectral fits (see below) and are certainly real. Therefore, we focus on these bright features in order to qualitatively investigate the overall variation in the X-ray line emission of Sgr A East. We present the Fe He$`\alpha `$ line ($`E`$ $``$ 6.6 keV) EW image of Sgr A East in Figure 2. The energy band selections for the line and continuum to generate this EW image are presented in Table 2. The Fe He$`\alpha `$ line EW is strongly enhanced within an $``$40<sup>′′</sup> diameter region near the center of the SNR. This Fe EW distribution is consistent with the central concentration of the Fe abundance in Sgr A East as reported with the previous Chandra and XMM-Newton observations (M02; S04). The high Fe EW at the center of the SNR is thus most likely caused by the enhanced Fe abundance there, displaying the distribution of Fe-rich stellar ejecta material. (Note: The bright X-ray emission at the position of Sgr A\* is featureless in the Fe EW image. This is reasonable for the continuum-dominated X-ray spectrum from Sgr A\* \[Baganoff et al. 2003\], which further supports the utility of our EW images.) In contrast to the central Fe-rich region, the northern plume region of the SNR does not show an enhancement in Fe, which suggests that it is produced by emission from shocked ISM having metal abundances close to the average value at the Galactic center. Unlike Fe, the EW maps of other elemental species S, Ar, and Ca do not show any significant spatial structure, and we do not explicitly present those EW images. The uniform EW distributions of other species are also consistent with the previously reported elemental abundance distributions across Sgr A East. ### 3.3 X-Ray Spectra We perform a spatially-resolved spectral analysis of Sgr A East utilizing high angular resolution ACIS data. For the spectral analysis of our CTI-corrected data, we use the response matrices appropriate for the spectral redistribution of the CCD, as generated by Townsley et al. (2002b). Because of various roll-angles of the individual ACIS observations and the moderate angular extension of the SNR ($``$3), Sgr A East has been detected on different ACIS-I chips depending on the roll-angles. Observations taken with deep exposures between 2002 February and 2002 June (ObsID 2943 $``$ 3665 in Table 1), however, had similar roll-angles, which represents $``$90% of the total exposure. Small sections of the SNR have thus been detected on a single CCD for the majority of the observations. For instance, although the bright, central Fe-rich region (i.e., “center” region in Figure 1) was detected on either the ACIS-I1 or the ACIS-I3 depending on the roll-angles, $``$91% of the total counts in the 1 $``$ 8 keV band were detected on the ACIS-I1. We thus use detector responses appropriate for the chip positions of the source region on the ACIS-I1 for the “center” region spectrum. We used the same method for other regional spectra. We binned the data to contain a minimum of 30 $``$ 50 counts per channel for the spectral fittings. Even though we corrected the data for the CTI, there remain some residual effects of small gain shifts, which are usually insignificant in the spectral analysis. The gain shift at $``$1% level may, however, confuse the spectral analysis of strong emission lines in a relatively narrow energy band such as Sgr A East spectrum containing strong Fe K lines between $`E`$ $``$ 6.4 keV $``$ 6.9 keV. We inspected any residual gain shifts with the ACIS-I calibration data (ObsID 61184) taken close to our deepest observations (2002 May 25). We found that there were $``$0.1% $``$ 1% gain shifts depending on the detector positions. We thus included a “gain fit” in the initial spectral fittings and then adjusted the gain according to the best-fit gain shift. The adjusted gain shifts are typically small ($``$0.6%), and are consistent with those from the calibration data. The background estimates for the spectral analysis of Sgr A East are difficult because X-ray background emission in the vicinity of Sgr A East shows complex spectral and spatial structure. We were particularly cautious about the background emission features from the neutral Fe K ($`E`$ $``$ 6.4 keV) and the H-like Fe K ($`E`$ $``$ 6.96 keV) lines, because line emission from these species is prevalent across the Galactic center region (Park et al. 2004a, ; Muno et al. 2004a, ). We considered several background regions which are source-free and represent the Fe line features. We eventually used the combined background emission from three regions which, we believe, adequately represent the “average” background spectrum for Sgr A East. The adopted background regions are presented in Figure 3. The spectral analysis and discussion in the following sections are based on this background selection. #### 3.3.1 Central Region We first extracted and fitted the spectrum of the $``$14<sup>′′</sup> $`\times `$ 24<sup>′′</sup> region containing the bright, Fe-rich emission at the core of the SNR (“center” in Figure 1 & Figure 2). This region selection was also intended to include the central region with the highest broadband intensity, while excluding the chip-gap. Even with the small angular size of $``$20<sup>′′</sup>, the “center” spectrum contains significant photon statistics of $``$7000 counts, and clearly shows the strong Fe He$`\alpha `$ line at $`E`$ $``$ 6.6 keV (Figure 4). Since M02 and S04 consistently found that thermal plasma in a collisional ionization equilibrium (CIE) can adequately describe the Sgr A East spectrum, we fitted the “center” spectrum with a CIE model (vmekal in XSPEC) absorbed by interstellar gas (we have also fitted the spectrum with a non-equilibrium ionization \[NEI\] model and the best-fit ionization timescale was indeed high \[$`n_et`$ $``$ 10<sup>12</sup> cm<sup>-3</sup> s\], indicating a CIE plasma condition). The fitted absorption column was large ($`N_H`$ $``$ 15 $`\times `$ 10<sup>22</sup> cm<sup>-2</sup>), which was consistent with the Galactic center location of the SNR. The electron temperature ($`kT`$ $``$ 2 keV) was consistent with that reported by M02. The high Fe abundance of a few times solar was also in agreement with the results of M02 and S04. Although a single temperature CIE model may describe the observed spectrum, we recognized a residual “hump” at $`E`$ $``$ 7 keV above the best-fit model, which appears to be a weak Fe Ly$`\alpha `$ line feature ($`E`$ $``$ 6.96 keV). In fact, the relatively poor fit ($`\chi _\nu ^2`$ $``$ 1.4) was primarily caused by this excess emission. Our attempts to remove these residuals by adopting background subtraction from various regions (including the same background regions used by M02 and S04) were not successful. We thus conclude that this residual emission originates from the SNR. The detection of the Fe Ly$`\alpha `$ line is indeed statistically significant ($``$7$`\sigma `$). This excess emission at $`E`$ $``$ 7 keV due to the H-like Fe line is important because an additional hot plasma component is required in order to fit this feature. The two-temperature CIE model can then improve the overall fit significantly ($`\chi _\nu ^2`$ $``$ 1.0; Figure 4). In this two-component fit, we varied S, Ca, Ar, and Fe abundances while keeping them in common for both components. All other species were fixed at solar abundance because the contribution from those elemental species in the fitted energy band is negligible. We also assumed the same foreground column for both components. The results of this two-component spectral fit are presented in Table 3 and Table 4. The fitted Fe abundance is high ($``$5.8 solar) while those for S ($``$0.7), Ar ($``$1.8), and Ca ($``$1.4) are relatively low. The electron temperatures from the best-fit model are $`kT`$ $``$ 1 keV and 5 keV, which are consistent with the results from S04. Although the electron temperature for the hard component, $`kT`$ $``$ 5 keV, appears to be unusually high for an SNR spectrum, such a temperature is required to properly describe the detected Fe Ly$`\alpha `$ line feature. This highly ionized H-like Fe line emission was not detected by the early Chandra observations with short exposures (M02). It was, however, suggested by the XMM-Newton data (S04), and is now confirmed by our deep Chandra observations. S04, with the XMM-Newton data, also reported detection of “neutral” Fe line emission ($`E`$ $``$ 6.4 keV) from Sgr A East. Although more prominent in the peripheral regions of the SNR, the neutral Fe K line was reportedly detected in the central regions of the SNR with XMM-Newton (photon flux $``$ 3 $`\times `$ 10<sup>-6</sup> photons cm<sup>-2</sup> s<sup>-1</sup> within 28<sup>′′</sup> radius; S04). Our deep, high angular resolution Chandra data do not require such a neutral Fe line feature in order to describe the observed spectrum. We place a 2$`\sigma `$ upper limit of 1.1 $`\times `$ 10<sup>-6</sup> photons cm<sup>-2</sup> s<sup>-1</sup> on the 6.4 keV Fe line photon flux. #### 3.3.2 “Plume” Regions We extract the spectrum of the plumelike feature from a region between the Fe-rich center and the cannonball (“plume” region in Figure 1 & Figure 2). The spectrum contains $``$5300 counts and can be fitted with a single-temperature ($`kT`$ $``$ 1.3 keV) CIE plasma model (Figure 5). In this fit, the elemental abundances for S, Ar, Ca, and Fe were varied freely, while abundances for other species were fixed at solar (Table 3 and Table 4). The Fe He$`\alpha `$ line is significantly weaker (e.g., EW $``$ 800 eV) than the “center” region spectrum (e.g., EW $``$ 2500 eV). The Fe abundance is also considerably lower than the central region, and is consistent with solar. These results indicate that the plume is emission primarily from a shocked ISM rather than from metal-rich ejecta. In contrast to the “center” region, hard excess emission at $`E`$ $``$ 7 keV due to the Fe Ly$`\alpha `$ line feature is not substantial in the plume region: i.e., the line detection is only marginal ($``$2.5$`\sigma `$), and the line flux is an order of magnitude lower than that from the “center” region. In fact, the statistical improvement with a two-temperature plasma model is insignificant (e.g., $`\chi _\nu ^2`$ $``$ 1.1 vs. 1.2). The fitted metal abundances also unchange when a two-temperature model is used. Although there appears to be a small contribution from the high temperature ($`kT`$ $``$ 5 keV) component, the overall spectrum from the plume region is thus dominated by the soft component (kT $``$ 1 keV). Therefore, we hereafter discuss the plume spectrum based on the single-temperature model, as presented in Table 3 and Table 4. We then extract the spectrum from a region between “center” and “plume” (“north” region in Figure 1 & Figure 2). This regional spectrum contains $``$7600 counts. The Fe He$`\alpha `$ line is substantially weaker than the “center” region, similar to the “plume” spectrum. On the other hand, as in the “center” region spectrum, hard excess emission at $`E`$ $`>`$ 7 keV was noticeable for the “north” region spectrum when fitted with a single-temperature model. We thus fit this regional spectrum with a two-temperature CIE plasma model in the same way as for the “center” region (Figure 6). The fitted parameters from the two-temperature model are presented in Table 3 and Table 4. The fitted parameters are generally consistent with those from the central region. It is, however, notable that the Fe abundance in the “north” region is significantly lower than that of the “center” region. As with the other regions, we found that the 6.4 keV line is not evident in the observed “north” region spectrum, with an upper limit to the line flux of 3.7 $`\times `$ 10<sup>-7</sup> photons cm<sup>-2</sup> s<sup>-1</sup> (2$`\sigma `$). ## 4 CXOGC J174545.5$``$285829: The Cannonball ### 4.1 X-Ray Morphology M02 speculated that a high-velocity neutron star associated with the SNR Sgr A East might exist at the “tip” of the plume. With deep observations, a hard pointlike source was indeed detected at the edge of the plume and designated as CXOGC J174545.5$``$285829 (the so-called cannonball) by Muno et al. (2003a). Figure 7 presents the soft and hard band images of the cannonball. In fact, at the edge of the plume, there are two pointlike sources separated by $``$3<sup>′′</sup>: the cannonball detected exclusively above $`E`$ $``$ 3 keV, and a soft source CXOGC J174545.2$``$285828 detected mostly below $`E`$ $``$ 2 keV (Figure 7). The soft source CXOGC J174545.2$``$285828 exhibits long-term variability (Muno et al. 2004b, ), and has an optical counterpart in the USNO-B1.0 catalog (#0610-0600649, m<sub>B</sub> = 15.27), both of which suggest that this soft source is a foreground star. Figure 7 reveals that the cannonball may not be truly pointlike, but is apparently extended: i.e., a cometary shape with a bright, pointlike “head” and a “tail” extending generally southward, which is roughly tracking back to Sgr A East. Although this source was detected $``$2$`.^{}`$3 off-axis, the apparent extended morphology is not caused by the point spread function (PSF). In Figure 8b, the one-dimensional source intensity distribution in the north-south direction is compared to that of the off-axis PSF at the same position. The extended nature of the cannonball toward the south is unambiguously revealed. In Figure 8a, we present a PSF-deconvolved source image, using a maximum-likelihood algorithm (Richardson, 1972; Lucy, 1974). For comparison, the deconvolved image of the soft source CXOGC J174545.2$``$285828 (the source in the right-hand side of the image) is also presented. The cannonball shows a significantly extended structure which confirms the cometary morphology suggested by the raw image and the projected intensity profile. The soft source, in contrast, appears to be resolved further into two pointlike sources. The photon statistics of the two sources, the cannonball and CXOGC J174545.2$``$285828, are similar ($``$1000 $``$ 1200 counts), and thus the morphological difference between the two is unlikely a statistical artifact. The angular extent of the cannonball’s tail is $``$3<sup>′′</sup>, which corresponds to a projected physical size of $``$0.1 pc (Hereafter, we use a distance of 8 kpc to the Galactic center \[Reid 1993\]). ### 4.2 X-Ray Spectrum The cannonball has been detected on various detector positions of the ACIS-I array because of the different roll-angles for the individual observations. These various detector positions include on-chip and chip-gap positions as well as different “rows” within the CCD. We thus use detector responses appropriate for each observation to average over all the observations for the spectral analysis of the combined data. For six observations (ObsIDs 242, 1561b, 2951, 2952, 2953, and 2954) in which the source was detected on an on-chip position, the source photons are relatively evenly spread over the ACIS-I0 (46% of the total) and the ACIS-I3 (54% of the total). We thus made a weighted sum (by detected counts) of the redistribution matrix function (RMF) files appropriate for the source locations of each CCD. The corresponding ancilliary response functions (ARF) were also averaged by weighting the counts detected in the individual observations. With the other five observations (ObsIDs 2943, 3663, 3392, 3393, and 3665), nearly all of the photons (90% of the total) were detected on the ACIS-I0. We thus use the RMF appropriate for the source detector positions on the I0 chip. The ARF files for these five observations were also averaged by weighting the counts detected in the individual observations. The source spectrum of the cannonball is extracted from an $``$1$`\stackrel{}{\mathrm{.}}`$8 radius circular region centered on the source position. We tested the background spectrum separately from three regions: i.e., an annulus of inner radius $``$5<sup>′′</sup> and outer radius $``$6$`\stackrel{}{\mathrm{.}}`$5 centered on the source position, a circular region of radius $``$11<sup>′′</sup> in the northeast side of the source, and a 10<sup>′′</sup> $`\times `$ 60<sup>′′</sup> box region along the CCD chip-gap (for ObsIDs 2943, 3663, 3392, 3393, and 3665) to the north of the source. The choice of these background regions resulted in no significant differences in the spectral fits. In the following discussion of the spectral analysis, we assume the rectangular background region along the chip-gap (Figure 1 & Figure 3), because this background region consistently represents the source “flux variation” caused by the on- and off-chip positions. We binned the source spectrum in order to contain 20 or more counts per channel. The spectrum can best be fitted with a power law (Figure 9) and the best-fit parameters are presented in Table 5. Because of the relatively low photon statistics ($``$1000 counts), the spectrum may also be fitted with thermal plasma models. The best-fit electron temperature is remarkably high ($`kT`$ $``$ 25 keV) with extremely low metal abundances. This high temperature might not be entirely unreasonable for sources like cataclysmic variables. Although unlikely, a thermal origin of the X-ray spectrum of the cannonball may thus not be completely ruled out. ## 5 DISCUSSION ### 5.1 The Cannonball: A High-Velocity Neutron Star? The spectral and morphological characteristics of the cannonball are very unusual among the sources detected within the ACIS FOV of the Galactic center: i.e., it is a bright source with a hard, continuum-dominated X-ray spectrum having no Fe line emission. This source also shows no long-term variability (Muno et al. 2004b, ). The cannonball is one of only two sources, out of $``$2300 Galactic center X-ray sources cataloged by Muno et al. (2003a), which share such unusual X-ray characteristics. We discuss this unique source in the following sections. The observed X-ray nature of the cannonball suggests an identification as a high-velocity neutron star in the Galactic center. Based on its proximity to Sgr A East, we therefore propose that the cannonball may be a candidate neutron star born from the SN explosion which also produced the SNR, Sgr A East. #### 5.1.1 Spectrum and Morphology The X-ray spectrum of the cannonball is best-fitted with a power law ($`\mathrm{\Gamma }`$ $``$ 1.6), which is typical for nonthermal synchrotron emission from a neutron star’s magnetosphere. The implied high foreground absorption is consistent with that of Sgr A East, in agreement with a Galactic center location nearby the SNR. Assuming a Galactic center location, the estimated X-ray luminosity (L<sub>X</sub> $``$ 3.1 $`\times `$ 10<sup>33</sup> ergs s<sup>-1</sup>) is also typical for a pulsar and/or pulsar wind nebula (PWN) (e.g., Becker & Aschenbach 2002). The deep Chandra images unambiguously reveal a cometary morphology of the cannonball. The tail is faint (only $`<`$100 photons), and the estimated physical size is $``$0.1 pc. This tail size is consistent with cometary tails detected from other Galactic high-velocity pulsars such as the Geminga pulsar (Caraveo et al., 2003), PSR B1853+01 (Petre et al., 2002), and PSR J1509$``$5850 (Sanwal et al. in preparation). The tail points roughly toward the south, consistent with the direction of the center of Sgr A East, given its limited photon statistics. The sky position of the source is also interesting: i.e., the cannonball is located, in projection, just interior to the radio shell boundary of Sgr A East while sitting on the “tip” of the X-ray plume. These X-ray morphologies are suggestive of a high-velocity neutron star, moving toward the north, for the origin of the cannonball. Considering an SNR age of $``$5000 $``$ 10000 yr (Mezger et al. 1989; Uchida et al. 1998; M02; S04) and the angular separation of $``$2 between the source and the “center” of the SNR, a velocity of $`v`$ $``$ 455 $``$ 912 km s<sup>-1</sup>/sin$`\beta `$, where $`\beta `$ is the angle between the line of sight and the actual traveling direction of the neutron star, is implied. This velocity range is in good agreement with that of typical high-velocity pulsars in the Galaxy (Cordes & Chernoff, 1998). We may make an independent estimate of the velocity of the candidate neutron star by assuming that the plume is a bow-shock produced as the high-velocity neutron star encounters the ISM. Based on the conical shape of the plume as seen by the “red” emission feature in Figure 1a, we estimate an opening angle of $`\theta `$ $``$ 53. The Mach number Ma = \[sin($`\theta `$/2)\]<sup>-1</sup> is thus $``$2.2. The velocity of the candidate neutron star is then $`v`$ = $`c_s`$ Ma = $``$880 km s<sup>-1</sup>. In this estimate, we assumed a sound speed of the ambient gas $`c_s`$ = ($`\gamma `$$`kT`$/$`\mu `$$`m_p`$)$`^{\frac{1}{2}}`$ $``$ 400 km s<sup>-1</sup> where $`\gamma `$ = 5/3 for a monatomic, adiabatic gas, $`\mu `$ = 1 for protons, $`m_p`$ is the proton mass, and $`kT`$ $``$ 1 keV plasma in the plume region. We note that, in the velocity estimation from the Mach number, we used the projected opening angle for the bow-shock. If the inclination angle of the bow-shock was significant, the actual, de-projected opening angle could have been smaller, thus higher velocity would have been derived. Nonetheless, the velocity estimates from the two methods are in plausible agreement. A third velocity estimate can be derived by assuming that this NS candidate produced a PWN which has reached pressure equilibrium with the ISM. Assuming that the X-ray spectrum of the cannonball is primarily from a PWN of the neutron star, we may derive the rotational spin-down energy loss $`\dot{E}`$ by using an empirical relationship between $`\dot{E}`$ and the power law photon index of the PWN, $$\mathrm{\Gamma }_{PWN}=2.360.021\dot{E}_{40}^{\frac{1}{2}},$$ where $`\dot{E}_{40}`$ is the spin-down power in units of 10<sup>40</sup> ergs s<sup>-1</sup> (Gotthelf, 2003). The best-fit photon index of $`\mathrm{\Gamma }_{PWN}`$ = 1.6 implies $`\dot{E}`$ $``$ 7.6 $`\times `$ 10<sup>36</sup> ergs s<sup>-1</sup>. We may assume a pressure balance between the ram pressure of the PWN, $`P_{PWN}`$ = $`\dot{E}`$/(4$`\pi `$cR<sup>2</sup>) where R is the PWN radius for a spherical geometry, and the thermal pressure of the plume region, $`P_{th}`$ = 2$`n_e`$$`kT`$. The best-fit volume emission measure ($`EM`$) from the plume region implies an electron density of $`n_e`$ $``$ 7.4 $`f^{\frac{1}{2}}`$ cm<sup>-3</sup>, where $`f`$ is the X-ray emitting volume filling factor (we assumed a half-conical volume with a circular base of radius $``$ 25<sup>′′</sup> and a height $``$ 50<sup>′′</sup> for the “plume” region). The best-fit thermal plasma model for the plume region then implies $`P_{th}`$ (= $`P_{PWN}`$) $``$ 3.1 $`\times `$ 10<sup>-8</sup> ergs cm<sup>-3</sup>. (The PWN radius is accordingly estimated to be R $``$ 2.5 $`\times `$ 10<sup>16</sup> cm, which corresponds to $``$0$`\stackrel{}{\mathrm{.}}`$2 at $`d`$ $``$ 8 kpc. This small R is consistent with the pointlike detection of the “head” of the source by the ACIS.) $`P_{PWN}`$ may be equivalent to the external ram pressure $`ϵ`$<sub>k</sub> = $`\rho `$$`v^2`$ ($`\rho `$ is the mass density of the ISM derived from the plume region spectrum). The velocity is then derived to be $`v`$ $``$ 550 km s<sup>-1</sup>. We note that there were a number of assumptions in these estimates, e.g., the pressure balance $`P_{PWN}`$ = $`P_{th}`$ = $`ϵ`$<sub>k</sub>, the $`\dot{E}`$ estimate from the best-fit photon index of the PWN spectrum, and the bow-shock origin for the plume. Considering various assumptions and embedded uncertainties in the above three approaches for the velocity estimates, the agreement among all three results are remarkable. The consistency among these independent estimates of the velocity thus lends all of them additional credibility. #### 5.1.2 Temporal Characteristics No evidence for long-term (observation by observation) variability of the cannonball was detected (Muno et al. 2004b, ). The proximity of this source to the chip-gap during the long series of observations makes it difficult to determine whether there is short-term (within individual observation periods) variability. With a 2$``$10 keV band flux of $``$1.9 $`\times `$ 10<sup>-13</sup> ergs cm<sup>-2</sup> s<sup>-1</sup>, the cannonball is one of the brightest sources detected in the ACIS FOV: i.e., the 5th brightest in terms of the photon flux, and the 18th brightest in the number of counts among $``$2300 cataloged Galactic center sources (Muno et al. 2003a, ). According to Muno et al. (2004b), $``$80% of the short-term variables have a maximum photon flux lower than the photon flux of the cannonball. Also, $``$94% of the long-term variables have a maximum flux lower than the cannonball (Muno et al. 2004b, ). For comparison, the nearby soft source CXOGC J174545.2$``$285828 with similar photon statistics was identified as a long-term variable (Muno et al. 2004b, ). The non-variability of the cannonball is thus unlikely a statistical artifact, but should correspond to the true nature of this object. The constant lightcurve with a long time-basis of $``$3 yr, particularly at the “high” flux level of this source, is remarkably unusual among the Galactic center sources. This non-variability over a few years indicates that the cannonball is unlikely a background AGN, and supports a neutron star identification. In fact, based on the logN-logS relation for the contribution from extragalactic sources in the Galactic center (Muno et al. 2003a, ), a low probability of $``$1.2 $`\times `$ 10<sup>-4</sup> for a detection of a background extragalactic source at the flux level of the cannonball within the “plume” region is implied. The discovery of a pulsar at the position of the cannonball would conclusively identify this source as a neutron star. Our deep Chandra observations, however, used the standard 3.2 s time frame which may not be adequate for a pulsar search. We, therefore, searched for pulsations in a 40-ks archival XMM-Newton/EPIC observation pointed at Sgr A\* (observation sequence number 0111350101). The EPIC/PN instrument has a relatively short frame time of 73.4 ms even in full-window mode, so that our pulsar search was sensitive down to periods of $``$147 ms. Because of the poor angular resolution of the EPIC instrument, the two sources, the cannonball and CXOGC J174545.2$``$285828, are not resolved. In order to discriminate contamination from the nearby soft source CXOGC J174545.2$``$285828, we extract the source photons only from the hard band (3 $``$ 10 keV). We then detect $``$460 counts from the cannonball. With these XMM-Newton data, we do not detect significant evidence of a pulsar. We place an upper limit of $``$40% on the pulsed-fraction between 5 $`\times `$ 10<sup>-5</sup> Hz and 6.8 Hz. If the cannonball is a pulsar associated with Sgr A East and thus has an age of $``$10000 yr, it may have a Vela pulsar-like periodicity of $``$100 ms. This pulsation period is beyond the detectability of the used XMM-Newton time resolution. The presence of a pulsar for this neutron star candidate cannot thus be ruled out yet. Follow-up X-ray and/or radio observations with high time resolution instruments would therefore be helpful for the pulsar search for this NS candidate. ### 5.2 Sgr A East: Fe Ejecta Mass The enhancement of strong Fe K line emission at the center of the SNR is remarkable. This Fe line emission is most likely from the Fe-rich stellar ejecta from the progenitor, which was heated by the reverse shock. Assuming that the Fe-rich central region represents the bulk of the total Fe mass ejected from the progenitor star, this Fe-rich material provides a useful opportunity to investigate the progenitor mass, and thus the SN explosion type of Sgr A East. In order to estimate the total Fe ejecta mass, we first consider a simple geometry in which all Fe ejecta material is concentrated within the central $``$40<sup>′′</sup> diameter region where the Fe EW is the highest (e.g., Fe EW $`>`$ 1400 eV in Figure 2). We estimate the electron density based on the measured $`EM`$ from the spectral fit of the “center” region. We use a cylindrical geometry with an elliptical cross-section (major and minor axis radii of 12<sup>′′</sup> and 7<sup>′′</sup>, respectively) for the X-ray emitting volume of the “center” region. A path-length of 1.55 pc was assumed, corresponding to $``$40<sup>′′</sup> angular extent of the Fe enhancement at the center of the SNR. The X-ray emitting volume of $`V`$ = 1.82 $`f`$ $`\times `$ 10<sup>55</sup> cm<sup>3</sup> is thus used for the density estimate. We, for simplicity, assume a “pure” Fe-ejecta case for the estimate of the electron density: i.e., all electrons originate from the ionized Fe. Considering the dominant ionization states presented by two-temperature plasma, we assume electron-Fe ion density ratio of $`n_e`$ = 24$`n_{Fe}`$ for the He-like Fe. We also use the best-fit Fe abundance of 5.8 $`\times `$ solar for each of the soft and the hard components. Considering the large difference in the plasma temperatures of the two components, they are unlikely “co-spatial” (e.g., S04). We thus assume that each component occupies a separate volume and maintains a pressure equilibrium: i.e., $`n_{es}`$T<sub>s</sub> = $`n_{eh}`$T<sub>h</sub> and $`f`$ = $`f_s`$ \+ $`f_h`$, where $`n_{es}`$ and $`n_{eh}`$ are electron densities, T<sub>s</sub> and T<sub>h</sub> are electron temperatures, and $`f_s`$ and $`f_h`$ are volume filling factors for the soft and the hard component, respectively. The best-fit electron temperatures ($`kT`$ = 1.05 keV and 5.28 keV) correspond to $`f_s`$ = 0.48$`f`$ and $`f_h`$ = 0.52$`f`$. The best-fit $`EM`$s then imply Fe ion densities of $`n_{Fes}`$ $``$ 0.096$`f^{\frac{1}{2}}`$ cm<sup>-3</sup> and $`n_{Feh}`$ $``$ 0.019$`f^{\frac{1}{2}}`$ cm<sup>-3</sup>, where $`n_{Fes}`$ and $`n_{Feh}`$ are Fe ion densities for the soft and the hard component, respectively. If we assume a spherical volume with a diameter of $``$1.55 pc and with a dominant isotope <sup>56</sup>Fe, the total Fe mass, M<sub>Fe</sub> = 56($`n_{Fes}`$ \+ $`n_{Feh}`$)$`m_pV`$, is then estimated to be M<sub>Fe</sub> $``$ 0.15$`f^{\frac{1}{2}}`$ M. Because of our “pure” Fe ejecta assumption, this Fe mass is an upper limit. The standard SN nucleosynthesis models indicate that M<sub>Fe</sub> is $``$0.5M $``$ 0.8M for the Type Ia SNe (e.g., Nomoto et al. 1997a). The Type II models show a wide range of Fe mass depending on the progenitor masses: e.g., M<sub>Fe</sub> $``$ 0.15M for a 13M $``$ 15M progenitor, and M<sub>Fe</sub> $``$ 0.05M $``$ 0.08M for a 18M $``$ 70M progenitor (e.g., Nomoto et al. 1997b). The estimated upper limit on the Fe mass for Sgr A East (M<sub>Fe</sub> $``$ 0.15M) is thus consistent with a Type II origin from a 13M $``$ 15M progenitor. If our Fe mass estimates were significantly affected by the ISM contribution, the progenitor mass could be larger. Although the total Fe mass, as estimated assuming a simple spherical geometry, provided a useful constraint on the progenitor mass, we may also entertain a contribution to the total Fe mass of the SNR from the region where Fe is marginally enhanced. Inclusion of this additional Fe mass may provide an even more conservative limit on the progenitor mass. The Fe EWs are moderately enhanced in an extended region just outside of the $``$40<sup>′′</sup> Fe core (i.e., the region between the Fe EW of 800 eV and 1400 eV contours in Figure 2). This region may be represented by a “spherical shell” with an outer diameter of $``$1 and an inner diameter of $``$40<sup>′′</sup>. The Fe abundance in this region appears to be $``$3. We thus consider the contribution from this moderately Fe-rich region to the total Fe mass of the SNR. We assume that the path-length through the outer shell region (foreground and background of the Fe-core in projection) is $``$50% of that for the inner Fe core for the adopted geometry. With the pure ejecta assumption, we consider that $`n_e`$ and $`n_{Fe}`$ are simply proportional to the Fe abundance ratio between the two regions (i.e., 5.8 for the inner core and 3 for the outer shell). The contribution from the shell region to the measured $`EM`$ is then $``$10%. We thus use the fitted $`EM`$s from the “center” region by assuming an $``$90% contribution from the inner Fe core ($``$1.55 pc diameter, corresponding to $``$40<sup>′′</sup> angular size at 8 kpc) and an $``$10% contribution from the outer shell ($``$0.39 pc thickness, corresponding $``$10<sup>′′</sup> angular size at 8 kpc). We also assume the same fractional volume filling factors ($`f_s`$ = 0.48$`f`$ and $`f_h`$ = 0.52$`f`$) as above. The derived total Fe densities are $`n_{Fe}`$ $``$ 0.107$`f^{\frac{1}{2}}`$ cm<sup>-3</sup> for the inner Fe core and $`n_{Fe}`$ $``$ 0.038$`f^{\frac{1}{2}}`$ cm<sup>-3</sup> for the outer shell. We then estimate the total Fe mass from the $``$1 diameter region, M<sub>Fe</sub> $``$ 0.27$`f^{\frac{1}{2}}`$ M, which is $``$80% larger than that derived merely from an $``$40<sup>′′</sup> diameter spherical core. This Fe mass is again a conservative upper limit because of the pure Fe ejecta assumption. We find that this upper limit of the total Fe mass of Sgr A East is still considerably lower than that expected from Type Ia models. Based on these derived Fe mass limits, we suggest that Sgr A East was produced likely by a core-collapse SN explosion of a massive progenitor star. A core-collapse origin for the SNR Sgr A East then supports the proposed SNR-NS association between Sgr A East and the cannonball. We note, however, that a Type Ia origin for Sgr A East might not be completely ruled out, yet. Considering the relatively old age of the SNR, one might speculate that a significant fraction of the Fe-rich ejecta has already been intermixed with the ambient ISM. The apparent lack of abundance enhancements from other high-Z elemental species may support this scenario. In such a case, the derived upper limits of Fe ejecta mass could be underestimates. If the amount of “missing” Fe ejecta mass was sufficiently large ($``$several times larger than the observed Fe ejecta mass), the “true” total Fe ejecta mass might have been large enough to be consistent with a Type Ia origin for Sgr A East. ### 5.3 Comments on Fe Lines The high angular resolution of the ACIS coupled with the deep exposure should be useful to investigate some controversial results between previous Chandra (M02) and XMM-Newton (S04) observations of Sgr A East: i.e., the XMM-Newton data reported line emission from highly-ionized H-like Fe and low-ionization state “neutral” Fe as well as the strong Fe He$`\alpha `$ line (S04), while the previous Chandra data showed only strong Fe He$`\alpha `$ line without H-like and neutral Fe lines (M02). These differences in the detected Fe line features resulted in discrepancies in the plasma temperatures ($`kT`$ $``$ 2 keV for the Chandra data vs. $`kT`$ $``$ 1 keV and 4 keV for the XMM-Newton data). Our data from the deep Chandra observations show evidence of Fe K line emission from the highly-ionized H-like Fe atoms, which confirms the results reported with the XMM-Newton data. The detected H-like Fe K line feature ($`E`$ $``$ 6.96 keV) requires an extremely high temperature plasma of $`kT`$ $``$ 5 keV in addition to the $`kT`$ $``$ 1 keV component, which also confirms the results obtained with the XMM-Newton. We thus attribute the discrepancies between M02 and S04 on the H-like Fe line features to the poor photon statistics of the previous Chandra data. The hard component plasma temperature is, however, unusually high for an SNR, and the interpretation of its origin is difficult. The peculiar interstellar environments in the Galactic center, such as unusually high magnetic fields, might be responsible for this high temperature. For instance, Muno et al. (2004a) reported the existence of very high temperature ($`kT`$ $``$ 8 keV) thermal plasma prevailing in the Galactic center regions. On the other hand, we detect no significant evidence for the neutral Fe line emission from Sgr A East, in contrast to the results from the XMM-Newton observation. It is notable that the sky position of Sgr A East is on the northeastern side of Sgr A\*, where the background 6.4 keV neutral Fe line emission (and other emission line features) of the Galactic center region prevails (Park et al. 2004a, ). The reported detection of the neutral Fe line with the XMM-Newton might have thus resulted from the inclusion of background diffuse emission and point sources in the spectrum of Sgr A East, neither of which could be properly accounted for, because of the poorer angular resolution of XMM-Newton. It is important to avoid confusion by the complex background line emission in the Galactic center for the spectral analysis of Sgr A East. The deep exposure and the superb angular resolution of our Chandra data allowed us to use small regions for the spectral analysis. For instance, our “center” region is an order of magnitude smaller in the source area than the central region used by S04, while achieving good photon statistics. We also utilized background regions which thoroughly represent all emission line features around Sgr A East. The detected point sources were not removed in the XMM-Newton data analysis (S04). Because faint point sources in the Galactic center regions are typically neutral Fe line sources (Muno et al. 2004b, ), the contribution from the unremoved point sources in the XMM-Newton data may also be responsible for the reported 6.4 keV line there (S04). In order to test these technical differences between S04 and the current work, we extracted the spectrum from a relatively large-size region ($``$28<sup>′′</sup> radius) around the center of Sgr A East without removing the detected point sources. We used the same background region as used in S04. We then made an $``$4$`\sigma `$ detection of the 6.4 keV line. The best-fit photon flux from the neutral Fe line was $``$2.5 $`\times `$ 10<sup>-6</sup> photons cm<sup>-2</sup> s<sup>-1</sup> which was consistent with that reported by S04. We thus conclude that the previously reported 6.4 keV line emission from Sgr A East with the XMM-Newton data (S04) was residual contamination from the Galactic center background emission. ## 6 SUMMARY AND CONCLUSIONS We presented the results from imaging and spectral analyses of the SNR, Sgr A East in the Galactic center, using deep Chandra observations. We confirm the central concentration of Fe-rich stellar ejecta in Sgr A East. The X-ray spectrum from the central regions of the SNR shows X-ray line emission from highly ionized He-like and H-like Fe. These Fe line features require multiple temperature thermal plasma with $`kT`$ $``$ 1 keV and 5 keV in order to properly describe the observed spectrum. The strong Fe lines imply overabundant Fe in the center of the SNR. On the other hand, the soft X-ray emission extending to the northern side of the SNR can be described with a single temperature ($`kT`$ $``$ 1.3 keV) thermal plasma with solar abundances. This northern “plume” is thus likely X-ray emission from the shocked ISM rather than metal-rich SN ejecta. A hard pointlike source (the so-called cannonball) detected at the northern edge of the plume, designated as CXOGC J174545.5$``$285829 in the Galactic center source catalog, shows remarkably unusual X-ray characteristics. The morphological, spectral, and temporal properties of this source are suggestive of an identification as a high-velocity neutron star. The estimated Fe ejecta mass of Sgr A East is consistent with a Type II SN for the origin of Sgr A East. Based on the suggested Type II origin for Sgr A East, the likely identification of the cannonball as a high-velocity neutron star, and the proximity between the cannonball and Sgr A East, we propose that these objects comprise an SNR-NS association in the Galactic center. We note, however, that follow-up pulsar searches with high time-resolution X-ray and/or radio observations at the position of the cannonball will be needed in order to make a conclusive identification of a neutron star for this object. This work has been supported in parts by NASA contract NAS8-39073, NAS8-01128 for the Chandra X-Ray Observatory. MPM is supported by the Hubble Fellowship program through grant number HS-HF-01164.01-A.
warning/0506/cond-mat0506183.html
ar5iv
text
# Simple mean-field theory for a zero-temperature Fermi gas at a Feshbach resonance ## Abstract We present a simple two-channel mean field theory for a zero-temperature two-component Fermi gas in the neighborhood of a Feshbach resonance. Our results agree with recent experiments on the bare-molecule fraction as a function of magnetic field \[Partridge et al., cond-mat/0505353\]. Even in this strongly-coupled gas of <sup>6</sup>Li, the experimental results depend on the structure of the molecules formed in the Feshbach resonance and, therefore, are not universal. Introduction.–Magnetoassociation creates a molecule from a pair of colliding atoms when one of the spins flips in the presence of a magnetic field tuned near a Feshbach resonance STW76 ; TIE93 . Nevertheless, the ultracold community as a whole has only reluctantly acknowledged the role of molecules in Feshbach-resonant interactions, clinging instead to an all-atom single-channel model solely characterized by a tunable $`s`$-wave scattering length KITP . In this Letter we promote a simple molecule-explicit two-channel mean-field theory to replace all-atom models. We find an excellent agreement with an experiment PAR05 that should present a severe test for theories of all ilk. Our approach is built around two species of fermionic atoms that magnetoassociate into bosonic “bare” molecules. In the experiments PAR05 , a laser spectroscopy probe is applied that detects bare molecules only. Below resonance, where thermodynamics favors molecules, the challenge for single-channel theories is to produce two types of molecules out of the same atoms, bare molecules seen by the probe and something else. Instead, our two-channel model simply states that bare molecules come “dressed” with atom pairs. Second, a typical all-atom theory rests on the notion of universality, whereby the interactions between atoms can be lumped into a scattering length and microscopic details of the physics do not matter. Our approach also mimics universal behavior on the atom side of the resonance for strong atom-molecule coupling. However, the experiments PAR05 cover a wide magnetic-field range on both sides of the resonance and, according to our results, certain molecular physics details are needed to satisfactorily match theory and the experiments. Universality does not hold despite strong interactions. Elementary as our mean-field theory is, it neatly bypasses two limitations of all-atom theories. Model.–As before JAV04 ; MAC05 , we study two fermionic species of atoms ($`c_𝐤`$ and $`c_𝐤`$) that magnetoassociate into bosonic molecules ($`b_𝐤`$) within a version of the Bose-Fermi model RAN85 used to describe superconductivity beyond the usual EAG69 weak-coupling limit. The corresponding microscopic Hamiltonian is $`{\displaystyle \frac{H}{\mathrm{}}}`$ $`=`$ $`{\displaystyle \underset{𝐤}{}}[(\frac{1}{2}ϵ_𝐤+\delta )b_𝐤^{}b_𝐤+ϵ_𝐤(c_𝐤^{}c_𝐤+c_𝐤^{}c_𝐤)]`$ (1) $`+{\displaystyle \underset{𝐤,𝐤^{}}{}}\kappa _{𝐤,𝐤^{}}(b_{𝐤+𝐤^{}}^{}c_𝐤c_𝐤^{}+c_𝐤^{}^{}c_𝐤^{}b_{𝐤+𝐤^{}}).`$ Here $`\mathrm{}ϵ_𝐤=\mathrm{}^2k^2/2m`$ is the kinetic energy for an atom with wave vector k, $`\delta `$ is the detuning controlled by the magnetic field ($`\delta >0`$ corresponds to an open dissociation channel for the molecules), and $`\kappa _{𝐤,𝐤^{}}`$ are the atom-molecule coupling coefficients for two atoms with momenta $`\mathrm{}𝐤`$ and $`\mathrm{}𝐤^{}`$. The key approximation is to treat the boson operators $`b_𝐤`$ and $`b_𝐤^{}`$ as classical conjugate variables. Then the hierarchy of the equations of motion for fermion correlation functions truncates exactly at the level of pair correlations. As a technical approximation, we only keep the zero momentum bosons, i.e., a uniform molecular condensate. We write fermion occupation numbers and pair correlations as a function of free-atom energy $`\mathrm{}ϵ\mathrm{}ϵ_𝐤`$ as $`P(ϵ)=c_𝐤^{}c_𝐤`$=$`c_𝐤^{}c_𝐤`$, $`C(ϵ)=c_𝐤c_𝐤`$, and, given the invariant atom number $`N`$ arising from the conserved quantity $`\widehat{N}=_𝐤(2b_𝐤^{}b_𝐤+c_𝐤^{}c_𝐤+c_𝐤^{}c_𝐤)`$, define the molecular amplitude $`\beta =\sqrt{2/N}b_0`$ such that $`\beta ^2`$ is the fraction of atoms converted into molecules. The result is the equations of motion JAV04 $`i\dot{C}(ϵ)`$ $`=`$ $`2ϵC(ϵ)+\frac{1}{\sqrt{2}}\mathrm{\Omega }\beta f(ϵ)[12P(ϵ)],`$ (2) $`i\dot{P}(ϵ)`$ $`=`$ $`\frac{1}{\sqrt{2}}\mathrm{\Omega }f(ϵ)[\beta C^{}(ϵ)\beta ^{}C(ϵ)],`$ (3) $`i\dot{\beta }`$ $`=`$ $`\delta \beta +{\displaystyle \frac{3\mathrm{\Omega }}{2\sqrt{2}ϵ_{F}^{}{}_{}{}^{3/2}}}{\displaystyle 𝑑ϵ\sqrt{ϵ}f(ϵ)C(ϵ)}.`$ (4) Here $`\mathrm{\Omega }=\sqrt{N}\kappa _{0,0}`$ is the atom-molecule Rabi frequency, and $`f(ϵ)`$ \[with $`f(0)=1]`$ conveys the energy dependence of atom-molecule coupling. Since the coupling coefficient $`\kappa _{0,0}`$ is inversely proportional to the square root of the quantization volume $`V`$, the Rabi frequency is proportional to the square root of the invariant density $`\rho =N/V`$. Finally, $`\mathrm{}ϵ_F=\mathrm{}^2(3\pi ^2\rho )^{2/3}/2m`$ is the usual Fermi energy for a two-component gas which, so far, arises simply because the sum over wave vectors is replaced by an integral over the frequencies. The objective is to find the steady state of Eqs. (2)-(4) at zero temperature. These equations admit solutions of the form $`C(ϵ;t)C(ϵ)e^{2i\mu t}`$, $`\beta (t)\beta e^{2i\mu t}`$ and $`P(ϵ;t)P(ϵ)`$, where $`\mathrm{}\mu `$ and $`2\mathrm{}\mu `$ are recognized as the chemical potentials for atoms and molecules. In fact, there are too many solutions. We therefore impose the condition of maximal pairing, whereby atoms only come in pairs of $`𝐤`$ and $`𝐤`$. This is technically the usual BCS ansatz, but here it is not so much an ansatz as a natural consequence of how a pair of atoms couples to a molecule. In mathematical terms we have $`|C(ϵ)|^2=P(ϵ)P^2(ϵ)`$. Equations (2) and (4) then give the counterpart of the gap equation in BCS theory, $$\delta 2\mu =\frac{3\mathrm{\Omega }^2}{8ϵ_F^{3/2}}𝑑ϵf^2(ϵ)\left(\frac{\sqrt{ϵ}}{\sqrt{(ϵ\mu )^2+\mathrm{\Delta }(ϵ)^2}}\frac{1}{\sqrt{ϵ}}\right),$$ (5) where $`\mathrm{\Delta }(ϵ)=f(ϵ)\beta \mathrm{\Omega }/\sqrt{2}`$ is the analog of the pairing gap. As we MAC02 ; JAV04 ; MAC05 and others have done in both boson KOK02 and fermion TIM01 systems, the term $`1/\sqrt{ϵ}`$ has been added into the integrand manually and the corresponding (possibly infinite) constant on the left-hand side has been absorbed into a “renormalization” of the detuning $`\delta `$. One more equation is needed, and is found from the conservation of atom number $`\widehat{N}`$ that holds for the mean-field equations (2)-(4) as well. We have $$\beta ^2+\frac{3}{4ϵ_F^{3/2}}𝑑ϵ\sqrt{ϵ}\left(1\frac{ϵ\mu }{\sqrt{(ϵ\mu )^2+\mathrm{\Delta }(ϵ)^2}}\right)=1.$$ (6) So far our approach deals with bare atoms and molecules that would emerge if the Feshbach-resonance interaction were suddenly turned off. In more practical terms, the sample breaks up into bare atoms and molecules if the magnetic field is switched quickly enough far enough away from the Feshbach resonance. The rate of change of the detuning that separates quick from slow is $`|\dot{\delta }|\mathrm{\Omega }^2`$ MAC02 ; JAV04 , and fast switching may be a tall order for the 834 G Feshbach resonance in <sup>6</sup>Li. Here, though, alternative methods to access the bare objects have been demonstrated. The resonance deals with spin-1/2 atoms, the dressed molecules are triplets, and the atoms have a net spin equal to zero. A measurement of the magnetic moment per atom JOC03 should FAL04 therefore only see the bare molecules. A laser spectroscopic probe that couples only to the triplet molecules likewise allows for a measurement of the bare-molecule fraction PAR05 . An alternative angle helps with the interpretation of the results. Let us initially take a single zero-momentum bare molecule. The available state space is spanned by the states with either one molecule at zero momentum and no atoms, $`b_0^{}|0`$, or no molecule and pairs of atoms $`𝐤`$ and $`𝐤`$, $`c_𝐤^{}c_𝐤^{}|0`$, where $`|0`$ is the particle vacuum. We are at liberty to write the state vector as $$|\psi (t)=\frac{1}{\sqrt{2}}\beta (t)b_0^{}|0+\frac{1}{\sqrt{N}}\underset{𝐤}{}C_𝐤(t)c_𝐤^{}c_𝐤^{}|0.$$ (7) Assuming self-consistently that the $`C`$ coefficients only depend on energy, $`C_𝐤(t)C(ϵ;t)`$, the Hamiltonian (1) gives the time dependent Schrödinger equations for the coefficients $`\beta `$ and $`C(ϵ)`$ that coincide with Eqs. (2) and (4), except that in the counterpart of Eq. (2) we have unity in lieu of the factor $`[12P(ϵ)]`$. Were it not for this factor that obviously reflects the Pauli exclusion principle, our mean-field theory would represent a collection of independent dressed molecules. A stationary state of the form (7) is a molecule “dressed” with atom pairs MAC02 ; JAV04 ; MAC05 ; STO04 ; FAL04 . With the same renormalization that was applied in Eq. (5), the dressed molecule is dissociated for $`\delta >0`$ and has one bound state for $`\delta <0`$. The latter is the bound state that molecular spectroscopy REG03 ; BAR05 detects in a dilute gas. The effect of the renormalization is to put the position of the Feshbach resonance in a dilute gas at the detuning $`\delta =0`$. The properties of the dressed molecule depend on the coupling function $`f(ϵ)`$, which may be obtained in principle from molecular structure calculations. Setting $`f(ϵ)1`$ corresponds to a contact interaction for the conversion between atoms and molecules. In reality, though, the interaction has a finite range. Corresponding to a nonzero spatial range, there is a range in energy; at high enough atom energies the coupling between two bare atoms and a bare molecule must vanish. As before JAV04 ; MAC02 , we use here the model in which we simply cut off the coupling at an energy $`\mathrm{}M`$ and write $`f(ϵ)=\theta (Mϵ),`$ where $`\theta `$ is the Heaviside unit step function. The binding energy $`2\mathrm{}|\mu |`$ of the dressed molecule for a negative detuning $`|\delta |`$ may be solved from the transcendental equation $$|\delta |2|\mu |\frac{3}{4}\sqrt{|\mu |}\sqrt{\overline{\omega }}\mathrm{arctan}\sqrt{\frac{M}{|\mu |}}=0,$$ (8) with $`\overline{\omega }=\mathrm{\Omega }^4/ϵ_F^3`$. While $`\mathrm{\Omega }`$ and $`ϵ_F`$ depend on the density, in the single-molecule parameter $`\overline{\omega }`$ the density dependence duly cancels. Given the binding energy, the bare-molecule fraction $`\beta ^2`$ in the dressed-molecule state vector is easily found. A substantial molecular component, say, $`\beta ^2>1/2`$ emerges in the contact interaction case $`M=\mathrm{}`$ for detunings $`\delta <(7\pi ^2/512)\overline{\omega }0.1\overline{\omega }`$ and in the case $`M\overline{\omega }`$ for detunings $`\delta <(3/4)\sqrt{M\overline{\omega }}`$. Until further notice all results are for the contact interaction, $`M=\mathrm{}`$. Main Features.–In the limit of weak coupling, $`\mathrm{\Omega }ϵ_F`$, the relevant demarkation points for the detuning $`\delta `$ are $`0`$ (Feshbach resonance) and $`2ϵ_F`$; 2 because it takes two atoms to make a molecule. For $`\delta <0`$ the system is a condensate of bare molecules, for $`\delta >2ϵ_F`$ a weak-coupling BCS superfluid. For $`0<\delta <2ϵ_F`$ we have an atom-molecule mixture in which all atoms from the Fermi sea above the energy $`\mathrm{}\delta /2`$ have been converted into bare molecules with this same energy, so that $`\beta ^2=1[\delta /(2ϵ_F)]^{3/2}`$. As has been noted before ZWI04 ; STO04 ; JAV04 ; WIL04 ; MAC05 , the Fermi sea of atoms blocks the decay of the molecules even if the molecules are above the dissociation threshold. In the intermediate case with $`\mathrm{\Omega }ϵ_F`$ numerical results show a continuous transformation from bare molecules to dressed molecules, to a nondescript mixture of atoms and molecules, and finally to a weak-coupling BCS as the detuning is varied from well below the Feshbach resonance to well above it. Typical <sup>40</sup>K experiments REG04 with $`\mathrm{\Omega }4ϵ_F`$ JAV04 ; MAC05 belong to this category. In the present approach the calculated fraction of bare molecules at the Feshbach resonance is 6%, and ca. 8% for non-zero temperature MAC05 ; STO04 , enough to cast ambiguity on the observation REG04 of fermionic condensation. For the 834 G Feshbach resonance in <sup>6</sup>Li, $`\mathrm{\Omega }ϵ_F`$. The coupling is now so strong that dissociation of above-threshold molecules is no longer Pauli blocked MAC05 , which may bring universal single-channel models into play. Since the term universality comes in many shades, we frame the issues somewhat along the lines of Refs. DIE04 ; HO04 ; DEP04 , in the order of increasing degree of universality. First, there are single-channel theories where atoms and their interactions are the building blocks, and explicit molecules are absent. The question is, do single-channel and two-channel descriptions give the same results? Second, can the relevant atom-atom or atom-molecule interactions be rolled into a single parameter, such as a scattering length, so that the results of either theory or experiments are independent of the microscopic physics of the interactions? Third, given the interaction parameter, will the system reach a unitarity limit that does not even depend on this parameter anymore as the interaction strength increases? Zero-temperature strong-coupling BCS theory is a single-channel model characterized entirely by the scattering length, and gives results that are independent of the scattering length when the scattering length tends to infinity. In this strong form of universality the BCS theory agrees with experiments carried out close to the Feshbach resonance OHA02 ; BOU02 ; GUP02 , although it may or may not be quantitatively accurate PER04 ; CAR03 . Moreover, as may be deduced easily from the arguments of Ref. DIE04 , in the case $`\mathrm{\Omega }ϵ_F`$ and uniformly for all detunings $`\delta >0`$, our two-channel mean-field theory and the standard BCS theory give the same results. The product of Fermi wave number and scattering length, $`k_Fa`$, is the dimensionless scaling parameter of the BCS theory. The translation into our model reads $`k_Fa3\pi \mathrm{\Omega }^2/8ϵ_F\delta `$. On the other hand, on the side of large negative detunings and in the limit of strong interaction, the two-channel theory scales with the dressed-molecule parameter $`\delta /\overline{\omega }`$. This parameter cannot be expressed in terms of $`k_Fa`$ and experimental constants such as atom mass, so that BCS and two-channel theories are not equivalent. As the normalization equation in the BCS theory is Eq. (6) sans the explicit $`\beta ^2`$, an increasing fraction of bare molecules signals an increasing discrepancy between single- and two-channel mean-field theories. Non-universal Experiments.–Our two-channel theory covers all parameter regions, and has different scaling properties on different sides of the resonance. It is therefore of particular interest that the Rice group has recently measured PAR05 the bare-molecule fraction of a Fermi gas over a wide range of magnetic fields across the resonance. Comparisons between theory and experiment are complicated by the fact that the experimental density varies substantially between data points taken at different magnetic fields. Density information can be recovered from a representation of the data in terms of the parameter $`k_Fa`$ PAR05 , but $`k_Fa`$ is not a scaling variable in our theory and we cannot use it unambiguously as the abscissa when plotting the calculated results. For these reasons we use the average of the Fermi energies deduced from the quoted values of $`k_Fa`$ as our parametrization of the density. Since the interaction-strength parameter scales weakly with density, $`\mathrm{\Omega }/ϵ_F\rho ^{1/6}`$, we still expect a qualitatively valid comparison with experiments. The Fermi energy here is given by $`ϵ_F=7.46\times 2\pi \mathrm{kHz}`$. The detuning in terms of magnetic field $`B`$ is $`\delta =\mathrm{\Delta }\mu (BB_0)/\mathrm{}`$, where $`\mathrm{\Delta }\mu =2\mu _B`$ is the difference in magnetic moments between a molecule and an atom pair, and $`B_0`$ is the position of the Feshbach resonance. The standard relation between atom-atom scattering length and magnetic field is $`a=a_b\left[1+\mathrm{\Delta }/(BB_0)\right]`$, where $`a_b`$ is the background scattering length and $`\mathrm{\Delta }`$ is the magnetic-field width of the Feshbach resonance. For a comparison, we need to modify the conversion of parameters between BCS and our approach to exclude the background scattering length, so that $`k_F(aa_b)=3\pi \mathrm{\Omega }^2/8\delta ϵ_F`$. The relation between Fermi energy and wave number, of course, is $`\mathrm{}ϵ_F=\mathrm{}^2k_F^2/(2m)`$. These observations, together with the known resonance parameters BAR05 , give the molecular parameter $`\overline{\omega }=3.34\times 2\pi \mathrm{THz}`$, enormously large compared with any other frequency scale in the problem. Finally, the Rabi frequency is given by $`\mathrm{\Omega }/ϵ_F=(\overline{\omega }/ϵ_F)^{1/4}=146`$, which indicates the strong-coupling regime. We plot in Fig. 1 the bare-molecule fraction $`\beta ^2`$ as a function of the magnetic field $`B`$ from the Rice data (circles) and from our calculations assuming a contact interaction with $`M=\mathrm{}`$ (dashed line). In the experiments the conversion into bare molecules extends a few hundred Gauss below the resonance, while the contact-interaction theory predicts a scale $`0.1\mathrm{}\overline{\omega }/|\mathrm{\Delta }\mu |10\mathrm{T}`$ and completely misses a qualitative trend in the experimental data. The remedy is to invoke a finite range for atom-molecule conversion. We vary the parameter $`M`$ until what is essentially the relative root-mean square error between the experiments and the theory is minimized at $`M=254\times 2\pi \mathrm{kHz}`$. The corresponding theory curve is plotted as the solid line. In view of the qualitative assumptions about the Fermi energy in the experiments and the simplistic model for the energy dependence of atom-molecule coupling, the agreement between theory and experiment is now excellent. The magnetic-moment data JOC03 on the bare-molecule fraction covers a much smaller dynamical range than optical spectroscopy, and only on the molecule side of the resonance. The corresponding theory in Ref. FAL04 is a dressed-molecule argument for $`\delta <0`$ that explicitly builds in the background scattering length $`a_b=1405a_0`$. Interestingly, when converted into a length via $`M=\mathrm{}/(m\mathrm{}^2)`$, our fitted cutoff $`M`$ gives $`\mathrm{}=1500a_0`$. Conclusions.–A finite cutoff ($`M`$) indicates that a contact interaction independent of the details of molecular physics is not sufficient to reach even a qualitative agreement with the experiments. The Rice data manifestly contradicts universality. Besides, experimental detection of a bare-molecule fraction $`0<\beta ^2<1`$ within a presumably molecular gas poses a severe challenge to any single-channel theory. A single-channel model can be retrofitted to include molecules as bound states of atom pairs, which would be the counterpart of our dressed molecules. But what, then, are the separate bare molecules? We have studied a simple two-channel mean-field theory for a zero-temperature Fermi gas at a Feshbach resonance. The model applies seamlessly across the resonance; no mixing or matching to join the BCS and BEC regimes is needed, or allowed. Our results agree with recent measurements of the fraction of bare molecules at various magnetic fields PAR05 over a dynamical range of five orders of magnitude. Even though the <sup>6</sup>Li Fermi gas is strongly coupled, the experimental results are not universal but reflect a finite range of atom-molecule coupling. This work is supported in part by NSF (PHY-0354599) and NASA (NAG3-2880). Randy Hulet generously provided both data in a digital form, and valuable advise.
warning/0506/astro-ph0506011.html
ar5iv
text
# Direct detection of exo-planets: GQ LupiProceedings ESO Workshop on The power of optical/IR interferometry ## 1 Introduction: GQ Lupi and its companion Since several years, we have been searching for sub-stellar companions around young (up to 100 Myrs) nearby (up to 150 pc) stars, both brown dwarfs and giant planets. Young sub-stellar objects are self-luminous due to ongoing contraction and accretion, so that young stars are good targets. We have found several brown dwarf companion candidates and confirmed three of them by both proper motion and spectroscopy: TWA-5 B, HR 7329 B, and GSC 8047 B, all being few Myr young late M-type brown dwarfs. With K-band imaging using VLT/NaCo, Subaru/CIAO, and HST/PC, we detected a 6 mag fainter object $`0.7^{\prime \prime }`$ west of the classical T Tauri star GQ Lup, which is a clearly co-moving companion ($`10\sigma `$), but orbital motion was not yet detectable. The NaCo K-band spectrum yielded $``$ L1-2 (M9-L4) as spectral type. At $`140\pm 50`$ pc distance (Lupus I cloud), it can be placed into the H-R diagram. For more details, see Neuhäuser et al. (2005). According to our own calculations following Wuchterl & Tscharnuter (2003), it has 1 to 3 M<sub>jup</sub>, but according to Baraffe et al. (2002) and Burrows et al. (1997) models, it is anywhere between 3 and 42 M<sub>jup</sub>. The latter models are not applicable for the young GQ Lup and its companion, while the former does take into account the formation and collapse; the Wuchterl & Tscharnuter (2003) convection model is calibrated on the Sun. ## 2 Comparison with model atmospheres We use the new so-called GAIA-dusty grid (Brott & Hauschildt et al., in preparation), which is an updated version of the models from Allard et al. (2001). The updates include improved molecular dissociation constants as well as more dust species and their opacities. In addition, the models were computed for a convective mixing length parameter of 1.5 times the pressure scale height H<sub>p</sub>. It uses spherical symmetry, which is most important for young and sub-stellar objects as GQ Lup A and b with low gravities. The Allard et al. (2001) AMES-dusty sometimes had problems at very low gravity. We compare our K-band spectrum (see Neuhäuser et al. 2005) with the GAIA grid for temperatures T$`{}_{\mathrm{eff}}{}^{}=2000`$ and 2900 K and for gravities of $`\mathrm{log}g=0,2`$, and 4 (g in cgs units). See figure 1 for the comparison. The model spectrum with T$`{}_{\mathrm{eff}}{}^{}=2000`$ K is much better than the hotter temperature (where there is no water vapour absorption in the blue part), again indicating an early L spectral type. For the gravities, $`\mathrm{log}g=2`$ fits best ($`\mathrm{log}g=4`$ is better than $`\mathrm{log}g=0`$). This gravity is fully consistent with the gravity-sensitive CO index measured in our spectrum to be $`0.862\pm 0.035`$, yielding $`\mathrm{log}g=2.5\pm 0.8`$ according to Gorlova et al. (2003). We conclude $`\mathrm{log}g2.0`$ to 3.3. The good fit indicates that the new spherically symmetric GAIA-dusty model is better for low gravities than the former AMES-dusty. It is important to note that the GAIA-dusty models are applicable: They are independant of age and stand-alone, even without interior models like the Baraffe et al. (2002) models. They are used, however, as outer boundary conditions by, e.g., Baraffe et al. (2002) to provide their evolution models with a more realistic description of how an interior model loses energy through the atmosphere. Models of the solar atmosphere usually used are also stand-alone and independant of the interior and age. ## 3 Discussion: Mass estimate for the GQ Lup companion At $`140\pm 50`$ pc distance, with T$`{}_{\mathrm{eff}}{}^{}=2050\pm 450`$ K and the flux of the companion (K = $`13.10\pm 0.15`$ mag), we can estimate its radius to be $`1.2\pm 0.5`$ R<sub>jup</sub>. With this radius and $`\mathrm{log}g2.0`$ to 3.3, we can estimate its mass to be $`1`$ M<sub>jup</sub> (for $`\mathrm{log}g4`$ and 2 R<sub>jup</sub>, its $`6`$ M<sub>jup</sub>). The Wuchterl & Tscharnuter (2003) calculations (Fig. 4 in Neuhäuser et al. 2005) indicate $`1`$ to 3 M<sub>jup</sub>, co-eval with the star at $`1`$ Myr. Mohanty et al. (2004a) measured gravities for isolated young brown dwarfs and free-floating planetary mass objects. Their coolest objects have spectral type M7.5 and gravities as low as $`\mathrm{log}g=3.125`$ (GG Tau Bb). This lead Mohanty et al. (2004b) to mass estimations as low as $`10`$ M<sub>jup</sub>. GQ Lup A is younger than the Mohanty et al. Upper Sco objects. Its companion is at least as late in spectral type, probably even cooler. An object younger and cooler must be lower in mass. The GQ Lup companion is fainter than the faintest Mohanty et al. object (USco 128, $`9`$ M<sub>jup</sub>), so that the mass estimate for the GQ Lup companion is $`8`$ M<sub>jup</sub>. According to Burrows et al. (1997) and Baraffe et al. (2002), the mass of the companion could be anywhere between 3 and 42 M<sub>jup</sub>, as given in Neuhäuser et al. (2005). However, these models are not valid at the young age of GQ Lup, so that they are not applicable. According to both Mohanty et al. (2004b) and Close et al. (2005), the Baraffe et al. (2002) models overestimate the masses of young sub-stellar objects below $`30`$ M<sub>jup</sub>, but underestimate them above $`40`$ M<sub>jup</sub>. This is consistent with our results. Hence, according to all valid estimations, the mass of the GQ Lup companion is $`1`$ to 8 M<sub>jup</sub>, i.e. significantly below $`13`$ M<sub>jup</sub>, hence almost certainly a planet imaged directly, to be called GQ Lup b. Acknowledgements. We would like to thank Subu Mohanty and Gibor Basri for very fruitfull discussion about GQ Lupi b. Especially, we would like to acknowledge Gibor Basri for pointing us to the Mohanty et al. work. ## Index
warning/0506/cond-mat0506520.html
ar5iv
text
# Diagnostics for the ground state phase of a spin-2 Bose-Einstein condensate ## I Introduction A Bose-Einstein condensate (BEC) in an optical trap Kurn exhibits a rich variety of spin-related phenomena, such as various magnetic phases Stenger and spin domain formation Miesner . The nuclear spin of $`{}_{}{}^{87}\mathrm{Rb}`$ and $`{}_{}{}^{23}\mathrm{Na}`$ is $`3/2`$, and combined with the spin $`1/2`$ of the outermost electron, the possible hyperfine spins of these atomic species are $`f=1`$ and $`f=2`$. The $`f=1`$ BEC was first realized at MIT with $`{}_{}{}^{23}\mathrm{Na}`$ Kurn , for which the antiferromagnetic behavior was observed Stenger . On the other hand, $`{}_{}{}^{87}\mathrm{Rb}`$ in the $`f=1`$ hyperfine state was found to be ferromagnetic Schmal ; Chang . The $`f=2`$ $`{}_{}{}^{23}\mathrm{Na}`$ condensate was also realized by the MIT group Gorlitz , whose ground state phase at zero magnetic field is predicted to be antiferromagnetic Ciobanu . However, the ground state phase and the spin dynamics of the $`f=2`$ $`{}_{}{}^{23}\mathrm{Na}`$ BEC have not been studied because of its very short lifetime (a few milliseconds), due to the fact that the energy of the $`f=2`$ state is higher than that of the $`f=1`$ state. The $`f=2`$ $`{}_{}{}^{87}\mathrm{Rb}`$ BEC also lies energetically higher than its $`f=1`$ counterpart, but its lifetime is much larger ($`100`$ ms) due to a fortuitous coincidence of the singlet and triplet scattering lengths Julienne . The coherent spin dynamics of the $`f=2`$ BEC have been observed Schmal ; Kuwamoto . The ground state of the $`f=2`$ BEC of $`{}_{}{}^{87}\mathrm{Rb}`$ is predicted to be very close to the phase boundary between the antiferromagnetic and cyclic phases Ciobanu ; Klausen . According to calculations performed by Klausen et alKlausen , the ground state phase of this BEC at zero magnetic field is barely in the antiferromagnetic phase. Recent experiments performed by the Hamburg group Schmal and by the Gakushuin group Kuwamoto appear to support this prediction. However, due to the experimental uncertainties, the possibility that the ground state phase is cyclic has not been excluded. In the Hamburg experiment Schmal , the $`m=\pm 2`$ mixture was shown to be stable, which is the ground state of the antiferromagnetic phase. However, a magnetic field of 340 mG was applied to the system in this experiment. Since the magnetic field lowers the energy of the $`m=\pm 2`$ states due to the quadratic Zeeman effect, the observed stability of the $`m=\pm 2`$ mixture may be due to the presence of the magnetic field. That there is no spatial separation between the $`m=\pm 2`$ states is a necessary condition for a BEC to be antiferromagnetic, but it is not sufficient. In the Gakushuin experiment Kuwamoto , the spin dynamics starting from the $`m=0`$ state were investigated for various values of the magnetic field. It was found that the population of the $`m=\pm 2`$ components after 70 ms evolution increased with a decrease of the magnetic field, which might imply that the $`f=2`$ BEC is antiferromagnetic at zero magnetic field. However, the initial $`m=0`$ state is a highly excited state and the resultant state after 70 ms is far above the ground state. Thus, the experimental observations do not exclude the possibility that the ground state phase at zero magnetic field is cyclic. In fact, the Hamburg group observed that the spin configuration of the cyclic phase has a long lifetime Schmal ; Sengstock . There are two experimental difficulties that hinder the determination of the ground state phase. One is the short lifetime of $``$ 100 ms of the upper hyperfine manifold of $`{}_{}{}^{87}\mathrm{Rb}`$, which makes the equilibrium spin state hard to reach. In order to realize the ground state phase, we must equilibrate the system at least for a few seconds, as in the $`f=1`$ case Chang ; Stenger . The other difficulty is the small energy scale ($``$ 0.1 nK Schmal ) of the spin-singlet-pair term whose sign determines the ground state phase. In order to determine the sign of the spin-singlet pair energy, the magnetic field must be reduced so that the quadratic Zeeman energy is less than the spin-singlet pair energy. However, in such a low magnetic field the system may be affected by stray ac magnetic fields Chang . To circumvent these difficulties, we propose a method to determine the value of the spin-singlet pair energy by spin exchange dynamics. We show that by controlling the initial population and phase in each magnetic sublevel, we can measure the value of the singlet-pair energy from the initial spin dynamics. This method therefore has the advantage that equilibrating the system, which takes a long time, is not necessary. Furthermore, the quadratic Zeeman energy is allowed to exceed the singlet-pair energy and extreme suppression of stray magnetic fields is not required. This paper is organized as follows. Section II formulates the mean field and Bogoliubov theories of a spin-2 BEC in the presence of the quadratic Zeeman effect. Section III studies the spin dynamics in a homogeneous system and proposes a method to determine the spin-singlet pair energy from the spin exchange dynamics. Section IV numerically verifies the proposed method for a trapped system with two-body loss. Section V discusses the initial spin preparation and Sec. VI concludes the paper. Detailed derivations of the analytic solutions are given in the Appendix. ## II Mean field and Bogoliubov analysis for a homogeneous system ### II.1 Formulation of the problem From rotational symmetry, the low-energy interaction between two atoms with hyperfine spin $`f`$ can be classified in terms of the total spin $`=0,2,\mathrm{},2f`$, and each scattering channel can be described by the corresponding $`s`$-wave scattering length $`a_{}`$ Ho . In the case of $`f=2`$ there are three scattering channels $`=0`$, $`2`$, $`4`$, and the interaction is described by $`(4\pi \mathrm{}^2/M)(a_0𝒫_0+a_2𝒫_2+a_4𝒫_4)\delta (𝒓_1𝒓_2)`$, where $`M`$ is the mass of the atom and $`𝒫_{}`$ projects the state of two colliding atoms into the state of total spin $``$. The interaction Hamiltonian can be rewritten Koashi ; Ueda as $`(c_0+c_1𝑺_1𝑺_2+c_2𝒫_0)\delta (𝒓_1𝒓_2)`$, where $`𝑺_1`$ and $`𝑺_2`$ are the spin operators for atoms 1 and 2 and the interaction coefficients are given by $`c_0`$ $`=`$ $`{\displaystyle \frac{4\pi \mathrm{}^2}{M}}{\displaystyle \frac{4a_2+3a_4}{7}},`$ (1a) $`c_1`$ $`=`$ $`{\displaystyle \frac{4\pi \mathrm{}^2}{M}}{\displaystyle \frac{a_4a_2}{7}},`$ (1b) $`c_2`$ $`=`$ $`{\displaystyle \frac{4\pi \mathrm{}^2}{M}}{\displaystyle \frac{7a_010a_2+3a_4}{7}}.`$ (1c) For the linear and quadratic Zeeman effects, the energy of the atom depends on the magnetic field $`B`$ as $`pm+qm^2`$, where $`p=\mu _\mathrm{B}B/2`$ and $`q=(\mu _\mathrm{B}B)^2/(4\mathrm{\Delta }_{\mathrm{hf}})`$ with $`\mu _\mathrm{B}`$ the Bohr magneton and $`\mathrm{\Delta }_{\mathrm{hf}}`$ the hyperfine splitting between the states $`f=1`$ and $`f=2`$ Pethick . We assume $`q<0`$ throughout this paper, which is the case for spin-2 $`{}_{}{}^{87}\mathrm{Rb}`$. The mean-field energy of the entire system is given by $`E`$ $`=`$ $`{\displaystyle }d𝒓[{\displaystyle \underset{m=2}{\overset{2}{}}}\psi _m^{}({\displaystyle \frac{\mathrm{}^2}{2M}}^2+V+pm+qm^2)\psi _m`$ (2) $`+{\displaystyle \frac{c_0}{2}}n_{\mathrm{tot}}^2+{\displaystyle \frac{c_1}{2}}𝑭^2+{\displaystyle \frac{c_2}{2}}|A_{00}|^2],`$ where $`m=2,\mathrm{},2`$ denotes the magnetic sublevels, $`V`$ is an external potential, $$n_{\mathrm{tot}}=\underset{m=2}{\overset{2}{}}|\psi _m|^2$$ (3) is the total particle-number density, $$𝑭=\underset{mm^{}}{}\psi _m^{}𝑺_{mm^{}}\psi _m^{}$$ (4) is the spin vector with $`5\times 5`$ spin-2 matrix $`𝑺`$, and $$A_{00}=\frac{1}{\sqrt{5}}\left(2\psi _2\psi _22\psi _1\psi _1+\psi _0^2\right)$$ (5) is the singlet-pair amplitude. Minimizing the action, $$S=𝑑t(i\mathrm{}\underset{m}{}\psi _m^{}_t\psi _mE),$$ (6) gives the multicomponent Gross-Pitaevskii (GP) equations for the spin-2 BEC as $`i\mathrm{}{\displaystyle \frac{\psi _{\pm 2}}{t}}`$ $`=`$ $`\left({\displaystyle \frac{\mathrm{}^2}{2M}}^2+V\pm 2p+4q\right)\psi _{\pm 2}+c_0n_{\mathrm{tot}}\psi _{\pm 2}`$ (7a) $`+c_1\left(F_{}\psi _{\pm 1}\pm 2F_z\psi _{\pm 2}\right)+{\displaystyle \frac{c_2}{\sqrt{5}}}A_{00}\psi _2^{},`$ $`i\mathrm{}{\displaystyle \frac{\psi _{\pm 1}}{t}}`$ $`=`$ $`\left({\displaystyle \frac{\mathrm{}^2}{2M}}^2+V\pm p+q\right)\psi _{\pm 1}+c_0n_{\mathrm{tot}}\psi _{\pm 1}`$ (7b) $`+c_1\left({\displaystyle \frac{\sqrt{6}}{2}}F_{}\psi _0+F_\pm \psi _{\pm 2}\pm F_z\psi _{\pm 1}\right)`$ $`{\displaystyle \frac{c_2}{\sqrt{5}}}A_{00}\psi _1^{},`$ $`i\mathrm{}{\displaystyle \frac{\psi _0}{t}}`$ $`=`$ $`\left({\displaystyle \frac{\mathrm{}^2}{2M}}^2+V\right)\psi _0+c_0n_{\mathrm{tot}}\psi _0`$ (7c) $`+{\displaystyle \frac{\sqrt{6}}{2}}c_1\left(F_{}\psi _1+F_+\psi _1\right)+{\displaystyle \frac{c_2}{\sqrt{5}}}A_{00}\psi _0^{},`$ where $`F_\pm =F_x\pm iF_y`$. We note that the linear Zeeman term only rotates the hyperfine spin about the $`z`$ axis, and hence the spin dynamics are essentially independent of the linear Zeeman term. In fact, setting $`\psi _me^{ipmt/\mathrm{}}\psi _m`$ and noting that the phase factor also arises in the $`F_\pm `$ terms as $`e^{ipt/\mathrm{}}F_\pm `$, we can completely eliminate the linear Zeeman terms in Eqs. (7a) and (7b). This property is due to the rotational symmetry of our system with respect to the $`z`$ axis, which results in the conservation of the projected angular momentum on the $`z`$ axis, $`F_z=𝑑𝒓_mm|\psi _m|^2`$. ### II.2 Ground states in a homogeneous system In an ultracold spinor BEC which is isolated from the environment, the total spin angular momentum in the direction of the magnetic field is conserved for a long time ($``$ 1 sec) Chang . We therefore minimize the energy of the system in the subspace of a given $`F_z=𝑑𝒓_mm|\psi _m|^2`$; we shall refer to the resulting minimized state as the “ground state”. For simplicity, we restrict ourselves to the subspace of $`F_z=0`$ in this paper, i.e., we seek the ground state in the subspace of $`F_z=0`$. We consider the uniform case by setting $`V=0`$ and $`\psi _m=\sqrt{n}\zeta _m`$, where the density $`n`$ is constant and $`\zeta _m`$ satisfies $`_m|\zeta _m|^2=1`$. The energy per atom is given by $$\epsilon \frac{E}{N}=q\underset{m=2}{\overset{2}{}}m^2|\zeta _m|^2+\frac{\stackrel{~}{c}_0}{2}+\frac{\stackrel{~}{c}_1}{2}𝒇^2+\frac{\stackrel{~}{c}_2}{2}|a_{00}|^2,$$ (8) where $`𝒇𝑭/n=_{mm^{}}\zeta _m^{}𝑺_{mm^{}}\zeta _m^{}`$, $`a_{00}=A_{00}/n=(2\zeta _2\zeta _22\zeta _1\zeta _1+\zeta _0^2)/\sqrt{5}`$, and $$\stackrel{~}{c}_ic_in$$ (9) for $`i=0`$, $`1`$, $`2`$. The linear Zeeman term is absent in Eq. (8) because $`F_z=0`$. In order to find the ground state, we compare the energies of the stationary states given in Ref. Sengstock . The energy of the antiferromagnetic state, $$𝜻_{\mathrm{AF}}=(e^{i\chi _2},0,0,0,e^{i\chi _2}),$$ (10) where $`\chi _m`$ is the phase of each component, is given by $$\epsilon _{\mathrm{AF}}=4q+\stackrel{~}{c}_0/2+\stackrel{~}{c}_2/10.$$ (11) This energy is independent of the phase factor $`e^{i\chi _m}`$. For the cyclic state, $$𝜻_\mathrm{C}=(\frac{1}{\sqrt{2}}e^{i\chi _2}\mathrm{sin}\theta ,0,e^{i\chi _0}\mathrm{cos}\theta ,0,\frac{1}{\sqrt{2}}e^{i\chi _2}\mathrm{sin}\theta ),$$ (12) the energy of the system takes the form $$\epsilon _\mathrm{C}=\frac{\stackrel{~}{c}_0}{2}+\frac{\stackrel{~}{c}_2}{10}\left|\mathrm{cos}^2\theta +e^{i\chi }\mathrm{sin}^2\theta \right|^2+4q\mathrm{sin}^2\theta ,$$ (13) where $`\chi \chi _2+\chi _22\chi _0`$. When $`\stackrel{~}{c}_2<10|q|`$, Eq. (13) is minimized by $`\theta =\pi /2`$, which corresponds to the antiferromagnetic state $`𝜻_{\mathrm{AF}}`$. When $`\stackrel{~}{c}_2>10|q|`$, Eq. (13) becomes minimal with $`\chi =\pi `$ and $$\mathrm{cos}^2\theta =\frac{1}{2}+\frac{5q}{\stackrel{~}{c}_2}.$$ (14) Substituting this into Eq. (13), we obtain $$\epsilon _\mathrm{C}=2q+\frac{\stackrel{~}{c}_0}{2}\frac{10q^2}{\stackrel{~}{c}_2}(\stackrel{~}{c}_2>10|q|).$$ (15) The states $`𝜻_{\mathrm{AF}}`$ and $`𝜻_\mathrm{C}`$ satisfy $`F_z(𝒓)=_mm|\psi _m(𝒓)|^2=0`$. Even if the local spin $`F_z(𝒓)`$ is nonzero, the total spin $`F_z`$ can be zero by the formation of the spatial spin structure Saito . For example, a staggered domain structure of the ferromagnetic state gives $`F_z=0`$. The kinetic energy at the domain wall can be neglected when the region of the wall is much smaller than the region of the domain. Hence, we also consider the case of $`F_z(𝒓)0`$, bearing in mind that the total spin is maintained as zero, $`F_z=0`$, by some structure formation. The energy of the ferromagnetic state, $`𝜻_\mathrm{F}=(e^{i\chi _2},0,0,0,0)`$ or $`(0,0,0,0,e^{i\chi _2})`$, is given by $$\epsilon _\mathrm{F}=4q+\stackrel{~}{c}_0/2+2\stackrel{~}{c}_1.$$ (16) We also consider a state found in Ref. Sengstock , $`𝜻_\mathrm{M}`$ $`=`$ $`(e^{i\chi _2}\mathrm{cos}\theta ,0,0,e^{i\chi _1}\mathrm{sin}\theta ,0)`$ (17) or $`(0,e^{i\chi _1}\mathrm{sin}\theta ,0,0,e^{i\chi _2}\mathrm{cos}\theta ).`$ The energy of this state is obtained to be $$\epsilon _\mathrm{M}=q(1+4\mathrm{cos}^2\theta )+\frac{\stackrel{~}{c}_0}{2}+\frac{\stackrel{~}{c}_1}{2}(3\mathrm{cos}^2\theta 1)^2.$$ (18) This is minimized by $`\mathrm{cos}^2\theta =1/3q/(3\stackrel{~}{c}_1)`$ for $`\stackrel{~}{c}_1>|q|/2`$, and the energy then becomes $$\epsilon _\mathrm{M}=2q+\frac{\stackrel{~}{c}_0}{2}\frac{q^2}{2\stackrel{~}{c}_1}(\stackrel{~}{c}_1>|q|/2).$$ (19) Comparing the energies $`\epsilon _{\mathrm{AF}}`$, $`\epsilon _\mathrm{F}`$, $`\epsilon _\mathrm{C}`$, and $`\epsilon _\mathrm{M}`$, we obtain the phase diagrams shown in Fig. 1. We have assumed here that one of the states $`𝜻_{\mathrm{AF}}`$, $`𝜻_\mathrm{C}`$, $`𝜻_\mathrm{F}`$, and $`𝜻_\mathrm{M}`$ is the ground state, based on the numerical results of Ref. Sengstock . We have confirmed that this assumption is correct using the Monte Carlo method. In the case of $`{}_{}{}^{87}\mathrm{Rb}`$, $`c_1`$ is positive and $`|c_2|`$ is at most of the same order of magnitude as $`c_1`$ Klausen . Hence, the ground state of the spin-2 $`{}_{}{}^{87}\mathrm{Rb}`$ BEC is antiferromagnetic or cyclic depending on whether $`\stackrel{~}{c}_2<10|q|`$ or $`\stackrel{~}{c}_2>10|q|`$. Therefore, in order to determine the sign of $`c_2`$ from the ground state, we must suppress the magnetic field so that the condition $`|q|<\stackrel{~}{c}_2/10`$ is met. In the experiment of the Hamburg group Schmal , a magnetic field of $`B=340`$ mG was applied, which corresponds to $`|q|/k_\mathrm{B}0.4`$ nK. The mean atomic density was $`n4\times 10^{14}\mathrm{cm}^3`$, and if $`|c_2|M/(4\pi \mathrm{}^2)a_\mathrm{B}`$, we have $`|\stackrel{~}{c}_2|/k_\mathrm{B}1.5`$ nK. Thus, $`\stackrel{~}{c}_210|q|`$, which is consistent with the observed stability of the antiferromagnetic state $`𝜻_{\mathrm{AF}}`$. In the Gakushuin experiment Kuwamoto , $`B>100`$ mG and $`n2.1\times 10^{14}\mathrm{cm}^3`$, corresponding to $`|q|0.03`$ nK and $`|\stackrel{~}{c}_2|/k_\mathrm{B}0.8`$ nK. Hence, the reported parameters of Ref. Kuwamoto belong to the cyclic phase. ### II.3 Stability of the stationary states The antiferromagnetic (or stretched) state, $$𝚿_{\mathrm{AF}}=\sqrt{\frac{n}{2}}(1,0,0,0,1),$$ (20) was found in experiments to be stable in the presence of a magnetic field Schmal . To examine this result, we will investigate the stability of this state using Bogoliubov analysis. Substituting Eq. (20) into Eqs. (7), we find that it evolves as $`e^{i\mu t/\mathrm{}}(e^{2ipt/\mathrm{}},0,0,0,e^{2ipt/\mathrm{}})`$ with $`\mu =4q+\stackrel{~}{c}_0+\stackrel{~}{c}_2/5`$, and hence Eq. (20) is stationary in the rotating frame $`e^{imp}\psi _m`$. We can therefore perform Bogoliubov analysis with respect to the state (20) in the rotating frame, which is described by the GP equations (7) without the linear Zeeman terms. We set the wave function as $`𝝍=e^{i\mu t/\mathrm{}}(𝚿_{\mathrm{AF}}+\mathit{\varphi })`$ and substitute it into the GP equations in the rotating frame. Taking the linear terms with respect to $`\mathit{\varphi }`$, we obtain $`i\mathrm{}{\displaystyle \frac{\varphi _{\pm 2}}{t}}`$ $`=`$ $`{\displaystyle \frac{\mathrm{}^2}{2M}}^2\varphi _{\pm 2}+\left({\displaystyle \frac{\stackrel{~}{c}_0}{2}}+2\stackrel{~}{c}_1\right)(\varphi _{\pm 2}+\varphi _{\pm 2}^{})`$ (21a) $`+\left({\displaystyle \frac{\stackrel{~}{c}_0}{2}}2\stackrel{~}{c}_1+{\displaystyle \frac{\stackrel{~}{c}_2}{5}}\right)(\widehat{\varphi }_2+\varphi _2^{}),`$ $`i\mathrm{}{\displaystyle \frac{\varphi _{\pm 1}}{t}}`$ $`=`$ $`\left({\displaystyle \frac{\mathrm{}^2}{2M}}^23q\right)\varphi _{\pm 1}`$ (21b) $`+\left(\stackrel{~}{c}_1{\displaystyle \frac{\stackrel{~}{c}_2}{5}}\right)(\varphi _{\pm 1}+\widehat{\varphi }_1^{}),`$ $`i\mathrm{}{\displaystyle \frac{\varphi _0}{t}}`$ $`=`$ $`\left({\displaystyle \frac{\mathrm{}^2}{2M}}^24q{\displaystyle \frac{\stackrel{~}{c}_2}{5}}\right)\varphi _0+{\displaystyle \frac{\stackrel{~}{c}_2}{5}}\varphi _0^{},`$ (21c) These equations can be reduced to the eigenvalue problem by expansion of $`\mathit{\varphi }`$ as $$\mathit{\varphi }(𝒓,t)=\underset{𝒌}{}\left(𝒖_𝒌e^{i(𝒌𝒓\omega _kt)}+𝒗_𝒌^{}e^{i(𝒌𝒓\omega _kt)}\right),$$ (22) and the eigenenergies are obtained as $`\left[\epsilon _k\left(\epsilon _k+2\stackrel{~}{c}_0+2\stackrel{~}{c}_2/5\right)\right]^{1/2},`$ (23a) $`\left[\epsilon _k\left(\epsilon _k+8\stackrel{~}{c}_12\stackrel{~}{c}_2/5\right)\right]^{1/2},`$ (23b) $`\left[(\epsilon _k3q)\left(\epsilon _k3q+2\stackrel{~}{c}_12\stackrel{~}{c}_2/5\right)\right]^{1/2},`$ (23c) $`\left[(\epsilon _k4q)\left(\epsilon _k4q2\stackrel{~}{c}_2/5\right)\right]^{1/2},`$ (23d) where $`\epsilon _k(\mathrm{}k)^2/(2M)`$. If the eigenenergy is complex, the corresponding mode is dynamically unstable against exponential growth. The eigenvectors $`𝒖_𝒌`$ and $`𝒗_𝒌`$ of the first mode (23a) are proportional to $`(1,0,0,0,1)`$, which is the same form as $`𝚿_{\mathrm{AF}}`$, indicating that a collapse occurs if $`2\stackrel{~}{c}_0+2\stackrel{~}{c}_2/5<0`$. The eigenvectors of the second mode (23b) are proportional to $`(1,0,0,0,1)`$. Since this eigenmode transfers the $`m=\pm 2`$ component to the $`m=2`$ component, excitation of the mode (23b) gives rise to exchange of atoms between the $`m=\pm 2`$ components. This implies that phase separation between the two components occurs for $`8\stackrel{~}{c}_12\stackrel{~}{c}_2/5<0`$. Therefore, the fact that no phase separation is observed in experiments Schmal only indicates that $`\stackrel{~}{c}_2<20\stackrel{~}{c}_1`$; it is not sufficient to conclude that $`\stackrel{~}{c}_2<0`$. The third mode (23c) is two-fold degenerate and the eigenvectors are proportional to $`(0,1,0,0,0)`$ and $`(0,0,0,1,0)`$. Therefore, when $`2\stackrel{~}{c}_12\stackrel{~}{c}_2/5<3q`$, the state (20) is dynamically unstable against the growth of the $`m=\pm 1`$ components. Similarly the last mode (23d) is proportional to $`(0,0,1,0,0)`$ and describes the growth of the $`m=0`$ component. Dynamical instability arises in this mode for $`\stackrel{~}{c}_2>10|q|`$, consistent with the fact that the ground state has the $`m=0`$ component for $`\stackrel{~}{c}_2>10|q|`$, as shown by Eq. (14). For later use, we also perform Bogoliubov analysis for the stationary state $`𝚿=(0,0,\sqrt{n},0,0)`$. The eigenenergies are obtained to be $`\left[\epsilon _k\left(\epsilon _k+2\stackrel{~}{c}_0+2\stackrel{~}{c}_2/5\right)\right]^{1/2},`$ (24a) $`\left[(\epsilon _k+q)\left(\epsilon _k+q+6\stackrel{~}{c}_12\stackrel{~}{c}_2/5\right)\right]^{1/2},`$ (24b) $`\left[(\epsilon _k+4q)\left(\epsilon _k+4q2\stackrel{~}{c}_2/5\right)\right]^{1/2}.`$ (24c) The eigenvector of the first mode (24a) is proportional to $`(0,0,1,0,0)𝚿`$, implying that the state $`𝚿`$ collapses if $`\stackrel{~}{2}c_0+2\stackrel{~}{c}_2/5<0`$. We note that this condition for the collapse of the BEC is the same as that for $`𝚿_{\mathrm{AF}}`$. The second and third modes, Eqs. (24b) and (24c), are both two-fold degenerate. The eigenvectors of the mode (24b) are proportional to $`(0,1,0,0,0)`$ and $`(0,0,0,1,0)`$, and therefore the dynamical instability in this mode increases the $`m=\pm 1`$ components. Similarly, the mode (24c) has eigenvectors proportional to $`(1,0,0,0,0)`$ and $`(0,0,0,0,1)`$. In contrast to the excitations for $`𝚿_{\mathrm{AF}}`$, there are always dynamically unstable modes in Eqs. (24b) and (24c) because of $`q<0`$, and therefore the $`m=0`$ state is always dynamically unstable for $`B>0`$ in a homogeneous system. ## III Spin dynamics in a homogeneous system ### III.1 Analytic solutions In order to determine the value of $`c_2`$ from the spin dynamics, we consider the spin dependent interaction that is sensitive only to the spin-singlet term. The elementary processes in the spin-singlet channel are $`0+00+0`$, $`0+01+(1)`$, $`0+02+(2)`$, and $`1+(1)2+(2)`$. On the other hand, the $`c_1`$ term consists of elementary processes such as $`m_1+m_2(m_1+1)+(m_21)`$. The spin independent interaction does not flip the spin. Thus, the elementary process appearing only in the $`c_2`$ term is $`0+02+(2)`$. We therefore focus on this process and consider the initial state of the form $$𝜻=(\frac{1}{\sqrt{2}}e^{i\chi _2}\mathrm{sin}\theta ,0,e^{i\chi _0}\mathrm{cos}\theta ,0,\frac{1}{\sqrt{2}}e^{i\chi _2}\mathrm{sin}\theta ).$$ (25) Since the overall phase and spin rotation about the $`z`$ axis do not affect the physics, the relevant phase appears only in the combination of $`\chi _2+\chi _22\chi _0`$. We therefore assume that $`\chi _0=\chi _2=0`$ and $`0\theta \pi /2`$ without loss of generality. If $`\zeta _{\pm 1}`$ are exactly zero in the initial state, we find from Eq. (7b) that $`\zeta _{\pm 1}(t)`$ are always zero within the mean field approximation. We therefore assume $`\zeta _{\pm 1}(t)=0`$ in the following analysis. In experiments Schmal ; Sengstock , this assumption holds at least for an initial period of $`100`$ ms. Although the $`m=\pm 1`$ components exponentially grow in the presence of the dynamical instability, the assumption $`\zeta _{\pm 1}(t)=0`$ is still valid in the early stage of time evolution. The GP equations then reduce to $`i\mathrm{}\dot{\zeta }_0`$ $`=`$ $`{\displaystyle \frac{\stackrel{~}{c}_2}{5}}\zeta _0^{}(2\zeta _2\zeta _2+\zeta _0^2),`$ (26a) $`i\mathrm{}\dot{\zeta }_{\pm 2}`$ $`=`$ $`{\displaystyle \frac{\stackrel{~}{c}_2}{5}}\zeta _2^{}(2\zeta _2\zeta _2+\zeta _0^2)+4q\zeta _{\pm 2}.`$ (26b) Differentiating Eq. (26a) with respect to time and using Eq. (26b), we can eliminate $`\zeta _{\pm 2}`$ to obtain $$\mathrm{}^2\ddot{\zeta }_0=\frac{2\stackrel{~}{c}_2}{5}\left(\epsilon 4q+8q|\zeta _0|^2\right)\zeta _0i\left(\frac{2\stackrel{~}{c}_2}{5}+8q\right)\mathrm{}\dot{\zeta }_0,$$ (27) where the energy per atom, $$\epsilon =\frac{\stackrel{~}{c}_2}{10}\left|2\zeta _2\zeta _2+\zeta _0^2\right|^2+4q\left(|\zeta _2|^2+|\zeta _2|^2\right),$$ (28) is a constant of motion. It is interesting to note that the form of Eq. (27) coincides with that describing a 1D BEC on a rotating ring if the time derivative is replaced with the spatial derivative Kanamoto . The solution is obtained (see Appendix for derivation) as $$\zeta _0(t)=e^{i\phi _0(t)}\sqrt{A_0\mathrm{sn}^2(\alpha t+\beta _0|\nu )+B_0},$$ (29) where $`\phi _0(t)`$, $`A_0`$, $`B_0`$, $`\alpha `$, $`\beta _0`$, and $`\nu `$ are given by Eqs. (70), (A), (61), (67), (68), and (63), respectively. The function sn is a Jacobian elliptic function Handbook . If we assume $`\chi _2=\pi `$ in the initial state, i.e., $$𝜻(0)=(\frac{1}{\sqrt{2}}\mathrm{sin}\theta _0,0,\mathrm{cos}\theta _0,0,\frac{1}{\sqrt{2}}\mathrm{sin}\theta _0),$$ (30) the constants in the solution (29) become $`\beta _0=0`$, $`B_0=\zeta _0^2(0)=\mathrm{cos}^2\theta _0`$, and $`A_0`$ $`=`$ $`{\displaystyle \frac{\stackrel{~}{c}_2}{10q}}\zeta _0^2(0)\left[1\zeta _0^2(0)\right],`$ (31) $`\alpha `$ $`=`$ $`\sqrt{{\displaystyle \frac{8q}{5\mathrm{}^2}}\left\{\stackrel{~}{c}_2\left[12\zeta _0^2(0)\right]+10q\right\}},`$ (32) $`\nu `$ $`=`$ $`{\displaystyle \frac{\stackrel{~}{c}_2^2\zeta _0^2(0)\left[1\zeta _0^2(0)\right]}{10q\left\{\stackrel{~}{c}_2\left[12\zeta _0^2(0)\right]+10q\right\}}}.`$ (33) The sign of $`A_0`$ is opposite to that of $`c_2`$ because $`q<0`$. The constants $`\alpha `$ and $`\nu `$ are real and positive if $`\stackrel{~}{c}_2\left[12\zeta _0^2(0)\right]+10q<0`$. ### III.2 Measurement of $`c_2`$ We suppose that the initial state (30) is prepared and the magnetic field satisfies $`|q|>|\stackrel{~}{c}_2|/10`$. Under this condition, we find that $`\alpha `$ and $`\nu `$ in Eqs. (32) and (33) are real and positive. Since the atomic density is typically $`n10^{14}\mathrm{cm}^3`$ in experiments Schmal ; Kuwamoto and $`|(7a_010a_2+3a_4)/7|`$ is of order of the Bohr radius at most Klausen , the condition $`|q|>|\stackrel{~}{c}_2|/10`$ requires $`B100`$ mG. At this magnetic field, the ground state in the subspace of $`F_z=0`$ is the stretched state (20) irrespective of the sign of $`c_2`$. The Jacobian elliptic function $`\mathrm{sn}(\alpha t|\nu )`$ initially increases for positive $`\alpha `$ and $`\nu `$, and hence the $`m=0`$ component $`|\zeta _0(t)|^2`$ in Eq. (29) initially decreases for $`c_2>0`$ and increases for $`c_2<0`$. Therefore, we can determine the sign of $`c_2`$ from the sign of the initial change in the spin population, which is the main result of this paper. This method is suitable for atomic species with short lifetimes, since only the initial stage of time evolution is needed for the determination of $`c_2`$. Figure 2 shows time evolution of the spin populations which are obtained by numerically solving Eq. (26). The results confirm the relation between the sign of $`c_2`$ and the sign of the initial change in the spin populations. In Fig. 2, the $`m=\pm 1`$ components are assumed to have small initial values ($`\zeta _{\pm 1}=\sqrt{0.001}`$) to assess imperfect population transfer in realistic situations. We find that the dynamical instability in the $`m=\pm 1`$ components have little effect on the initial dynamics of the $`m=0`$, $`\pm 2`$ components. In Fig. 2, $`\nu 0.01`$. For $`0\nu 1`$, the Jacobian elliptic function can be approximated by the trigonometric function as Handbook $$\mathrm{sn}(\alpha t|\nu )=\mathrm{sin}\alpha t+O(\nu ),$$ (34) and the spin evolution is then approximated by $`|\zeta _0(t)|^2`$ $``$ $`\mathrm{cos}^2\theta _0+A_0\mathrm{sin}^2\alpha t,`$ (35a) $`|\zeta _{\pm 2}(t)|^2`$ $``$ $`{\displaystyle \frac{1}{2}}\left(\mathrm{sin}^2\theta _0A_0\mathrm{sin}^2\alpha t\right).`$ (35b) The amplitude of this oscillation is $`A_0c_2`$, showing explicitly that whether the spin population initially increases or decreases depends on the sign of $`c_2`$. We can also determine the magnitude of $`c_2`$ by measuring the amplitude of the oscillations. In Fig. 2, Eq. (35) is plotted as dashed curves, which are in good agreement with the numerical solutions (solid curves) in the early stages of the time evolution. The differences between the numerical and analytic curves in Fig. 2 arise mainly from the growth of the $`m=\pm 1`$ components in the numerical result due to the dynamical instability. We find that the growth of the $`m=\pm 1`$ components in Fig. 2 (b) is larger than that in Fig. 2 (a). This can qualitatively be understood from the Bogoliubov energy (24b), which describes the growth of the $`m=\pm 1`$ components from the state $`𝚿=(0,0,\sqrt{n},0,0)`$. The imaginary part of Eq. (24b) at $`\epsilon _k=0`$, $`[q(q+6\stackrel{~}{c}_12\stackrel{~}{c}_2/5)]^{1/2}`$, is larger for $`c_2<0`$ than for $`c_2>0`$ since $`q<0`$ and $`q+6\stackrel{~}{c}_1>0`$ in the present case. We note that the relative phase $`\chi \chi _2+\chi _22\chi _0=\pi `$ in the initial state (25) is crucial for determining the value of $`c_2`$ by the above method. For example, when the initial state is given by Eq. (25) with $`\chi =0`$, or without loss of generality, $$𝜻=(\frac{1}{\sqrt{2}}\mathrm{sin}\theta _0,0,\mathrm{cos}\theta _0,0,\frac{1}{\sqrt{2}}\mathrm{sin}\theta _0),$$ (36) the time evolution becomes Eq. (29), where $`A_0`$, $`\nu `$, and $`\alpha `$ are given by Eqs. (A), (84), and (85), respectively (see Appendix for details). These constants satisfy $`c_2A_0>0`$ and $`\nu <0`$, and $`\alpha `$ is real. Using the relation Handbook $$\mathrm{sn}(\alpha t||\nu |)=\frac{1}{\sqrt{1+|\nu |}}\mathrm{sd}\left(\sqrt{1+|\nu |}\alpha t\right|\frac{|\nu |}{1+|\nu |})$$ (37) and noting that the Jacobian elliptic function $`\mathrm{sd}=\mathrm{sn}/\mathrm{dn}`$ initially increases, we find that $`|\zeta _0(t)|^2`$ initially decreases for $`c_2<0`$ and increases for $`c_2>0`$. This behavior is opposite to that in Fig. 2 and therefore control of the relative phase $`\chi `$ in the initial state is crucial for determining the sign of $`c_2`$. Figure 3 shows the initial evolution of $`|\zeta _0|^2`$ with $`\chi `$ varying from $`0`$ to $`\pi `$, which indicates that the phase of the oscillation depends on $`\chi `$. ### III.3 Dynamics in the cyclic phase If $`|q|<\stackrel{~}{c}_2/10`$, the stretched state (20) is dynamically unstable against growth of the $`m=0`$ component, as shown by Eq. (23d). This is because the ground state is in the cyclic phase given by Eq. (12) if $`|q|<\stackrel{~}{c}_2/10`$ and $`\stackrel{~}{c}_2<20\stackrel{~}{c}_1`$. We can therefore conclude that $`\stackrel{~}{c}_210|q|>0`$ and hence $`c_2>0`$, if we observe the growth of the $`m=0`$ component from the initial stretched state. Suppose that $`\mathrm{cos}^2\theta _01`$ in the initial state given by Eq. (30) and $`|q|<\stackrel{~}{c}_2/10`$, then $`\nu `$ is negative and $`\alpha `$ is pure imaginary from Eqs. (32) and (33). In this case, the Jacobian elliptic function can be rewritten as $$\mathrm{sn}(i|\alpha |t||\nu |)=\frac{i}{\sqrt{1+|\nu |}}\mathrm{sd}(\sqrt{1+|\nu |}|\alpha |t|\frac{1}{1+|\nu |}).$$ (38) Since $`|\nu |1`$ and hence $`1/(1+|\nu |)1`$, we can make the approximation Handbook $$\mathrm{sd}(u|(1+|\nu |)^1)\frac{\mathrm{sinh}u+\frac{|\nu |}{4}(\mathrm{sinh}u\mathrm{cosh}uu)\mathrm{sech}u}{1+\frac{|\nu |}{4}(\mathrm{sinh}u\mathrm{cosh}uu)\mathrm{tanh}u},$$ (39) where $`u|\alpha |t`$. When $`u1`$, Eq. (39) reduces to $$\mathrm{sd}(u|(1+|\nu |)^1)\frac{1}{\sqrt{|\nu |}}\mathrm{sech}(|\alpha |t\frac{1}{2}\mathrm{ln}\frac{16}{|\nu |}).$$ (40) Equation (29) is thus approximated to give $$|\zeta _0(t)|^2\frac{A_0}{|\nu |}\mathrm{sech}^2\left(|\alpha |t\frac{1}{2}\mathrm{ln}\frac{16}{|\nu |}\right)+\mathrm{cos}^2\theta _0.$$ (41) The dashed curve in Fig. 4 shows Eq. (41). A comparison of this with the corresponding numerical result (solid curve) shows that Eq. (41) is in good agreement with the numerical result, with a small discrepancy arising mainly from the growth of the $`m=\pm 1`$ components in the numerical result due to the dynamical instability. The $`m=0`$ component reaches a maximum at $`t\mathrm{ln}(16/|\nu |)/(2|\alpha |)`$ and the maximum value of $`|\zeta _0(t)|^2`$ is given by $`A_0/|\nu |1+10q/\stackrel{~}{c}_2`$. We note that a characteristic time scale of the dynamics in Fig. 4 is $`100`$ ms, which is much larger than the $`10`$ ms seen in Fig. 2. This is because the former time scale is set by the dynamical instability, while the latter is set by the time evolution from a non-stationary state. This method can thus determine the value of $`c_2`$, provided that (i) $`c_2`$ is positive, (ii) the magnetic field is suppressed so that $`|q|<\stackrel{~}{c}_2/10`$, and (iii) the time scale of the dynamics is much faster than the lifetime of the $`f=2`$ state. If $`c_2`$ is negative, the present method cannot determine the value of $`c_2`$. In the case of Fig. 4 ($`|q|<\stackrel{~}{c}_2/10`$), the $`m=0`$ population initially increases for $`c_2>0`$, which is opposite to the case of Fig. 2 ($`|q|>|\stackrel{~}{c}_2|/10`$). Nevertheless, the method to determine the sign of $`c_2`$ in Sec. III.2 is still applicable, since we can easily find whether we are in a region of $`|q|>\stackrel{~}{c}_2/10`$ or $`|q|<\stackrel{~}{c}_2/10`$. If the amplitude of the initial oscillation monotonically decreases without changing sign with an increase in the magnetic field, we can conclude that $`|q|>\stackrel{~}{c}_2/10`$. ## IV Spin dynamics in the trapped system We investigate the spin dynamics in a trap potential to verify the results in Sec. III in a realistic situation. We use an axisymmetric trap potential with radial and axial frequencies $`(\omega _{},\omega _z)=2\pi \times (237,21)`$ Hz, which was used in the experiment of Ref. Kuwamoto . The two-body loss rate was found to be roughly independent of the magnetic field and therefore it is considered to be insensitive to the spin-mixing dynamics Kuwamoto . We therefore take into account the two-body loss by adding the term $`i\mathrm{}K_2|\psi _m|^2\psi _m/2`$ to the right-hand side of Eq. (7) with $`K_2=1.4\times 10^{13}\mathrm{cm}^3\mathrm{s}^1`$. We prepare the ground state $`\psi _g`$ for the $`m=2`$ state by the imaginary-time propagating method Dalfovo and transfer it to each spin component as $`\psi _0=\sqrt{0.298}\psi _g`$, $`\psi _{\pm 1}=\sqrt{0.001}\psi _g`$, and $`\psi _2=\psi _2=\sqrt{0.35}\psi _g`$. We use this state as an initial state, where the initial number of atoms is assumed to be $`N(t=0)=4.7\times 10^5`$ Kuwamoto . The time evolution of each spin population obtained by numerically solving the GP equation (7) with the two-body loss terms is shown in Fig. 5. We find that the aforementioned dependence of the initial spin evolution on the sign of $`c_2`$ is valid in the trapped system, i.e., the quantity $`|\zeta _0|^2=𝑑𝒓|\psi _0|^2/N`$ first decreases for $`c_2>0`$ and increases for $`c_2<0`$. The growth in $`|\psi _{\pm 1}|^2`$ due to the dynamical instability and the two-body loss does not affect this dependence. Thus, we have confirmed that the sign of $`c_2`$ can be determined from the spin dynamics in a realistic situation. We find that the growth of the $`m=\pm 1`$ components is larger for $`c_2<0`$ than for $`c_2>0`$, which is similar to the homogeneous case shown in Fig. 2. The insets in Fig. 5 show the column density distribution of each spin component at $`t=100`$ ms. The initial distribution is almost preserved at $`t=100`$ ms and no spin domains are formed. ## V Preparation of the initial spin state As shown in Fig. 3, our method to determine the sign of $`c_2`$ hinges on the relative phase $`\chi _2+\chi _22\chi _0`$ in the initial state (25). Therefore we must control the phase experimentally. In the experiments of Refs. Schmal ; Kuwamoto , the rapid-adiabatic-passage technique Mewes was employed to prepare the initial state. This method can control the phase in the initial state in principle, if the relative phases between the applied rf fields are controlled. We propose here a new method to prepare the initial state using the Raman transition. We consider the Raman transition, for example, between the $`f=2`$ and $`f^{}=2`$ $`D_1`$ states by circularly polarized photons, as illustrated in Fig. 6. The initial state is assumed to be the $`m=0`$ lower state. Hence, the relevant states are the $`m=0`$, $`\pm 2`$ lower states and the $`m=\pm 1`$ upper states, whose amplitudes are denoted by $`\zeta _0`$, $`\zeta _{\pm 2}`$, and $`\zeta _{\pm 1}^{}`$, respectively. The equations of motion are given by $`i\mathrm{}\dot{\zeta }_{\pm 2}`$ $`=`$ $`(\pm 2p+4q)\zeta _{\pm 2}{\displaystyle \frac{1}{\sqrt{6}}}g_{}^{}e^{i\omega _{}t}\zeta _{\pm 1}^{},`$ (42a) $`i\mathrm{}\dot{\zeta }_{\pm 1}^{}`$ $`=`$ $`(E_{D_1}\pm p^{}+q^{})\zeta _{\pm 1}^{}{\displaystyle \frac{1}{\sqrt{6}}}g_{}e^{i\omega _{}t}\zeta _{\pm 2}`$ (42b) $`\pm {\displaystyle \frac{1}{2}}g_\pm e^{i\omega _\pm t}\zeta _0,`$ $`i\mathrm{}\dot{\zeta }_0`$ $`=`$ $`{\displaystyle \frac{1}{2}}g_+^{}e^{i\omega _+t}\zeta _1^{}{\displaystyle \frac{1}{2}}g_{}^{}e^{i\omega _{}t}\zeta _1^{},`$ (42c) where $`g_\pm `$ and $`\omega _\pm `$ are the coupling constants and frequencies of the $`\sigma _\pm `$ photons, $`E_{D_1}`$ is the transition energy of the $`D_1`$ line, and $`p^{}`$ and $`q^{}`$ are the linear and quadratic Zeeman energies for the upper hyperfine manifold. The coupling constants $`g_\pm `$ are given by the product of the amplitude of the circularly polarized light and the dipole matrix element between the $`s`$ and $`p`$ states. Using the new variables $`\stackrel{~}{\zeta }_{\pm 2}=e^{\pm 2ipt}\zeta _{\pm 2}`$ and $`\stackrel{~}{\zeta }_{\pm 1}^{}=e^{i\omega _\pm t}\zeta _{\pm 1}`$, the coefficients on the right-hand side of Eq. (42) become time independent. Assuming $`\omega _\pm =E_{D_1}\mathrm{\Delta }\pm p+q^{}`$ and $`\mathrm{\Delta }|pp^{}|`$, and eliminating $`\zeta _{\pm 1}^{}`$ from Eq. (42), we obtain $`i\mathrm{}\dot{\stackrel{~}{\zeta }}_{\pm 2}`$ $`=`$ $`\left(4q{\displaystyle \frac{|g_{}|^2}{6\mathrm{\Delta }}}\right)\stackrel{~}{\zeta }_{\pm 2}+i\mathrm{}\mathrm{\Omega }\zeta _0,`$ (43a) $`i\mathrm{}\dot{\stackrel{~}{\zeta }}_0`$ $`=`$ $`{\displaystyle \frac{1}{4\mathrm{\Delta }}}\left(|g_+|^2+|g_{}|^2\right)\zeta _0i\mathrm{}(\mathrm{\Omega }^{}\stackrel{~}{\zeta }_2\mathrm{\Omega }\stackrel{~}{\zeta }_2),`$ where $`i\mathrm{}\mathrm{\Omega }=g_+g_{}^{}/(2\sqrt{6}\mathrm{\Delta })`$. If the amplitudes of the $`\sigma _\pm `$ fields are the same, i.e., $`|g_+|=|g_{}|`$, and if the magnetic field is applied so that the condition $`q=|g_+|^2/(12\mathrm{\Delta })`$ is met, the coefficients of the first terms on the right-hand side of Eq. (43) become equal, i.e., $`4q|g_{}|^2/(6\mathrm{\Delta })=(|g_+|^2+|g_{}|^2)/(4\mathrm{\Delta })=6q`$. The solutions are thus obtained as $`\zeta _2`$ $`=`$ $`e^{i(2p+6q)t/\mathrm{}}{\displaystyle \frac{\mathrm{\Omega }}{\sqrt{2|\mathrm{\Omega }|^2}}}\mathrm{sin}\sqrt{2|\mathrm{\Omega }|^2}t,`$ (44a) $`\zeta _0`$ $`=`$ $`e^{i6qt}\mathrm{cos}\sqrt{2|\mathrm{\Omega }|^2}t,`$ (44b) $`\zeta _2`$ $`=`$ $`e^{i(2p+6q)t/\mathrm{}}{\displaystyle \frac{\mathrm{\Omega }^{}}{\sqrt{2|\mathrm{\Omega }|^2}}}\mathrm{sin}\sqrt{2|\mathrm{\Omega }|^2}t.`$ (44c) We note that the phase difference is always $`\chi =\chi _2+\chi _22\chi _0=\pi `$, which is independent of the phase of the circularly polarized light and the applied magnetic field. This method is thus suitable for preparing the initial state (30). We can check whether the prepared state has the correct relative phase $`\chi =\pi `$. Rotating the spin state (25) by $`\pi /2`$ around the axis on the $`x`$-$`y`$ plane as $`\overline{𝜻}=e^{i\pi S_\varphi /2}𝜻`$ with $`S_\varphi =S_x\mathrm{cos}\varphi +S_y\mathrm{sin}\varphi `$, we find that the population of each component becomes $`|\overline{\zeta }_2|^2`$ $`=`$ $`|\overline{\zeta }_2|^2={\displaystyle \frac{1}{8}}(3\mathrm{cos}^2\theta \sqrt{3}\mathrm{sin}2\theta \mathrm{cos}{\displaystyle \frac{\chi }{2}}\mathrm{cos}\chi ^{}`$ (45a) $`+\mathrm{sin}^2\theta \mathrm{cos}^2\chi ^{}),`$ $`|\overline{\zeta }_1|^2`$ $`=`$ $`|\overline{\zeta }_1|^2={\displaystyle \frac{1}{2}}\mathrm{sin}^2\theta \mathrm{sin}^2\chi ^{},`$ (45b) $`|\overline{\zeta }_0|^2`$ $`=`$ $`{\displaystyle \frac{1}{4}}(\mathrm{cos}^2\theta +\sqrt{3}\mathrm{sin}2\theta \mathrm{cos}{\displaystyle \frac{\chi }{2}}\mathrm{cos}\chi ^{}`$ (45c) $`+3\mathrm{sin}^2\theta \mathrm{cos}^2\chi ^{}),`$ where $`\chi ^{}=2\varphi +(\chi _2\chi _2)/2`$. Let us consider the quantity $$\overline{f}_z^2=2(4|\overline{\zeta }_2|^2+|\overline{\zeta }_1|^2)=1+2\mathrm{cos}^2\theta \sqrt{3}\mathrm{sin}2\theta \mathrm{cos}\frac{\chi }{2}\mathrm{cos}\chi ^{},$$ (46) which can be measured by the spin-dependent non-destructive imaging reported in Ref. Higbie . We note that if $`\chi =\pi `$, Eq. (46) is independent of $`\chi ^{}`$ and is constant, while if $`\chi \pi `$, Eq. (46) oscillates, since the phase $`\chi _2\chi _2`$ rotates in the magnetic field. We can therefore confirm that the desired initial state (30) is obtained if the result of the polarization-dependent phase-contrast imaging Higbie is constant in time. ## VI Conclusions We have proposed a method to experimentally determine the sign of the singlet-pair energy of a spin-2 BEC from spin exchange dynamics, which determines whether the $`f=2`$ $`{}_{}{}^{87}\mathrm{Rb}`$ BEC at zero magnetic field is antiferromagnetic or cyclic. We have obtained analytic solutions of the multicomponent GP equations and found that if we prepare the initial state in the magnetic sublevels $`m=0`$, $`\pm 2`$ with the appropriate relative phase relationship, we can determine the sign of $`c_2`$ from the sign of the initial change in the spin populations. For example, in the case of the initial state in Eq. (25), we can conclude that $`c_2`$ is positive or negative if the $`m=0`$ component first decreases or increases. We can also determine the magnitude of $`c_2`$ from the amplitude of the oscillation in the spin populations. Since only the initial evolution for a period of $`10`$ ms is needed, we can use this method in the presence of atom loss by inelastic collisions and even in the presence of dynamical instabilities. We have numerically confirmed that this method is applicable for a trapped system with a realistic two-body loss coefficient $`K_2`$. The required condition for the magnetic field is $`|q|>\stackrel{~}{c}_2/10`$, so our method works even when the quadratic Zeeman energy exceeds the singlet-pair energy. The initial spin state in which the populations and relative phases are controlled as Eq. (25) can be prepared by, e.g., the Raman technique as shown in Sec. V. We have shown that the coherent spin dynamics serve as an efficient probe for investigating spinor properties especially when the atomic species have a short lifetime and therefore the equilibrium state cannot be reached. The coherent spin dynamics may also reveal other equilibrium spinor properties as well as nonequilibrium spinor properties that cannot be obtained from the static equilibrium. ###### Acknowledgements. We would like to thank T. Kuwamoto and T. Hirano for fruitful discussions. This work was supported by Grant-in-Aids for Scientific Research (Grant No. 17740263 and No. 15340129) and by a 21st Century COE program at Tokyo Tech “Nanometer-Scale Quantum Physics,” from the Ministry of Education, Culture, Sports, Science and Technology of Japan. ## Appendix A Derivation of the solution (29) In this appendix, we solve Eq. (26) for the initial condition (25) with $`\chi _0=\chi _2=0`$ and $`0\theta \pi /2`$, and derive the solution (29). Defining a new variable $$z_0(t)e^{i(\stackrel{~}{c}_2/5+4q)t/\mathrm{}}\zeta _0(t),$$ (47) we can rewrite Eq. (27) as $$\ddot{z}_0=Pz_0+Q|z_0|^2z_0,$$ (48) where $`P`$ $`=`$ $`{\displaystyle \frac{1}{\mathrm{}^2}}\left({\displaystyle \frac{2\stackrel{~}{c}_2}{5}}\epsilon _s{\displaystyle \frac{\stackrel{~}{c}_2^2}{25}}16q^2{\displaystyle \frac{16\stackrel{~}{c}_2q}{5}}\right),`$ (49) $`Q`$ $`=`$ $`{\displaystyle \frac{16\stackrel{~}{c}_2q}{5\mathrm{}^2}},`$ (50) with the spin-dependent energy $$\epsilon _s=\frac{\stackrel{~}{c}_2}{10}|2\zeta _2\zeta _2+\zeta _0^2|^2+4q(1|\zeta _0|^2).$$ (51) The initial conditions for $`z_0`$ are $`z_0(0)=\zeta _0(0)=\mathrm{cos}\theta _0`$ and $`\dot{z}_0(0)`$ $`=`$ $`{\displaystyle \frac{i}{\mathrm{}}}\left({\displaystyle \frac{\stackrel{~}{c}_2}{5}}+4q\right)\zeta _0(0)`$ (52) $`{\displaystyle \frac{i\stackrel{~}{c}_2}{5\mathrm{}}}\zeta _0(0)\left[2e^{i\chi _2}\zeta _2(0)^2+\zeta _0^2(0)\right],`$ where we have used Eq. (26a). We write the complex variable as $`z_0(t)R(t)e^{i\varphi (t)}`$ with real functions $`R(t)`$ and $`\varphi (t)`$, where $`\varphi (0)=0`$ and $`z_0(0)=R(0)`$. Substituting this into Eq. (48) and taking the real and imaginary parts, we obtain $`\ddot{R}PRQR^3R\dot{\varphi }^2`$ $`=`$ $`0,`$ (53) $`2\dot{R}\dot{\varphi }+R\ddot{\varphi }`$ $`=`$ $`0.`$ (54) These equations can be integrated to give $`\dot{R}^2`$ $`=`$ $`PR^2+{\displaystyle \frac{Q}{2}}R^4+C_R{\displaystyle \frac{C_\varphi ^2}{R^2}},`$ (55) $`\dot{\varphi }`$ $`=`$ $`{\displaystyle \frac{C_\varphi }{R^2}},`$ (56) where $`C_R`$ and $`C_\varphi `$ are constants of integration. The time derivative of $`z_0`$ at $`t=0`$, $`\dot{z}_0(0)=\dot{R}(0)+i\dot{\varphi }(0)R(0)`$, gives $`\mathrm{Re}\dot{z}_0(0)=\dot{R}(0)`$ and $`\mathrm{Im}\dot{z}_0(0)=\dot{\varphi }(0)R(0)`$. The constant $`C_\varphi `$ is then written as $$C_\varphi =\dot{\varphi }(0)R^2(0)=z_0(0)\mathrm{Im}\dot{z}_0(0).$$ (57) The constant $`C_R`$ is similarly obtained as $$C_R=|\dot{z}_0(0)|^2Pz_0^2(0)\frac{Q}{2}z_0^4(0).$$ (58) We introduce a new variable $`\rho `$ through $$R^2=A_0\rho ^2+B_0,$$ (59) where $`A_0`$ and $`B_0`$ are constants. Substituting Eq. (59) into Eq. (55), we obtain $`A_0^2\dot{\rho }^2`$ $`=`$ $`{\displaystyle \frac{Q}{2}}A_0^3\rho ^4+\left({\displaystyle \frac{3Q}{2}}B_0+P\right)A_0^2\rho ^2`$ $`+\left({\displaystyle \frac{3Q}{2}}B_0^2+2PB_0+C_R\right)A_0`$ $`+\left({\displaystyle \frac{Q}{2}}B_0^3+PB_0^2+C_RB_0C_\varphi ^2\right)\rho ^2.`$ We determine $`B_0`$ so that the $`\rho ^2`$ term in Eq. (A) vanishes: $$\frac{Q}{2}B_0^3+PB_0^2+C_RB_0C_\varphi ^2=0.$$ (61) The right-hand side (rhs) of Eq. (A) then becomes quartic with respect to $`\rho `$. We determine $`A_0`$ so that the rhs of Eq. (A) becomes proportional to $`(1\rho ^2)(1\nu \rho ^2)`$; the result is $`A_0`$ $`=`$ $`{\displaystyle \frac{1}{2Q}}[3QB_02P`$ $`\pm \sqrt{(3QB_02P)(QB_0+2P)8QC_R}]`$ with $$\nu =\frac{QA_0}{QA_0+2P+3QB_0}.$$ (63) Equation (A) thus becomes $$\dot{\rho }^2=\frac{QA_0}{2\nu }(1\rho ^2)(1\nu \rho ^2).$$ (64) This equation can be integrated to give $$_{\rho (0)}^{\rho (t)}\frac{d\rho }{\sqrt{(1\rho ^2)(1\nu \rho ^2)}}=\sqrt{\frac{QA_0}{2\nu }}t.$$ (65) We rewrite this equation as $$_0^{\rho (t)}\frac{d\rho }{\sqrt{(1\rho ^2)(1\nu \rho ^2)}}=\alpha t+\beta _0,$$ (66) where $`\alpha `$ $`=`$ $`\sqrt{{\displaystyle \frac{QA_0}{2\nu }}},`$ (67) $`\beta _0`$ $`=`$ $`{\displaystyle _0^{\rho (0)}}{\displaystyle \frac{d\rho }{\sqrt{(1\rho ^2)(1\nu \rho ^2)}}}.`$ (68) The left-hand side of Eq. (66) is an elliptic integral of the first kind and hence $`\rho (t)=\mathrm{sn}(\alpha t+\beta _0|\nu )`$ Handbook . Thus, $$|\zeta _0(t)|=R(t)=\sqrt{A_0\mathrm{sn}^2(\alpha t+\beta _0|\nu )+B_0}.$$ (69) From Eqs. (47) and (56), the phase $`\phi _0`$ of $`\zeta _0`$ is given by $$\phi _0(t)=\frac{1}{\mathrm{}}\left(\frac{\stackrel{~}{c}_2}{5}+4q\right)t+_0^t𝑑\tau \frac{C_\varphi }{A_0\mathrm{sn}^2(\alpha t+\beta _0|\nu )+B_0}.$$ (70) Similarly, we can obtain the solution for $`\zeta _{\pm 2}(t)`$. From the form of Eq. (26) and the initial condition $`\zeta _2(0)=e^{i\chi _2}\zeta _2(0)`$, we find that the relation $`\zeta _2(t)=e^{i\chi _2}\zeta _2(t)`$ always holds. Hence, we consider only $`\zeta _2(t)`$. Eliminating $`\zeta _0(t)`$ from Eq. (26), we have the equation of motion for $`\zeta _2(t)`$ as $$\mathrm{}^2\ddot{\zeta }_2=\left[\frac{2\stackrel{~}{c}_2}{5}(\epsilon _s+4q)16q^2\frac{32\stackrel{~}{c}_2q}{5}|\zeta _2|^2\right]\zeta _2i\frac{2\stackrel{~}{c}_2}{5}\mathrm{}\dot{\zeta }_2.$$ (71) The new variable $`z_2(t)=e^{i\stackrel{~}{c}_2t/(5\mathrm{})}\zeta _2(t)`$ reduces Eq. (71) to $$\ddot{z}_2=P^{}z_2+Q^{}|z_2|^2z_2,$$ (72) where $`P^{}=(2\stackrel{~}{c}_2\epsilon _s/5\stackrel{~}{c}_2^2/2516q^2+8\stackrel{~}{c}_2q/5)/\mathrm{}^2`$ and $`Q^{}=32\stackrel{~}{c}_2q/(5\mathrm{}^2)`$. The initial conditions for $`z_2`$ are $`z_2(0)=\zeta _2(0)=\mathrm{sin}\theta _0/\sqrt{2}`$ and $$\dot{z}_2(0)=\frac{i\stackrel{~}{c}_2}{5\mathrm{}}\zeta _2(0)\left[12\zeta _2^2(0)e^{i\chi _2}\zeta _0^2(0)\right]\frac{4iq}{\mathrm{}}\zeta _2(0).$$ (73) Equation (72) has the same form as Eq. (48), in which $`P`$ and $`Q`$ are replaced by $`P^{}`$ and $`Q^{}`$, and therefore the derivation from Eq. (53) to Eq. (70) can similarly be applied. We consider the case of $`\chi _2=\pi `$ in the initial condition given in Eq. (25). In this case, Eq. (52) is pure imaginary, and Eq. (61) can be solved to give $`B_0=z_0^2(0)=\zeta _0^2(0)`$. Equation (A) then reduces to $$A_0=\{\begin{array}{c}12\zeta _0^2(0)+\frac{10q}{\stackrel{~}{c}_2},\hfill \\ \frac{\stackrel{~}{c}_2}{10q}\zeta _0^2(0)\left[1\zeta _0^2(0)\right],\hfill \end{array}$$ (74) where the upper and lower expressions correspond to the plus and minus signs of $`\pm `$ in Eq. (A). Using this $`A_0`$, $`\nu `$ and $`\alpha `$ become $`\nu `$ $`=`$ $`\{\begin{array}{c}{\displaystyle \frac{10q\left\{\stackrel{~}{c}_2\left[12\zeta _0^2(0)\right]+10q\right\}}{\stackrel{~}{c}_2^2\zeta _0^2(0)\left[1\zeta _0^2(0)\right]}},\hfill \\ {\displaystyle \frac{\stackrel{~}{c}_2^2\zeta _0^2(0)\left[1\zeta _0^2(0)\right]}{10q\left\{\stackrel{~}{c}_2\left[12\zeta _0^2(0)\right]+10q\right\}}},\hfill \end{array}`$ (77) $`\alpha `$ $`=`$ $`\{\begin{array}{c}{\displaystyle \frac{2|\stackrel{~}{c}_2|}{5\mathrm{}}}\sqrt{\zeta _0^2(0)\left[1\zeta _0^2(0)\right]},\hfill \\ \sqrt{{\displaystyle \frac{8q}{5\mathrm{}^2}}\left\{\stackrel{~}{c}_2\left[12\zeta _0^2(0)\right]+10q\right\}}.\hfill \end{array}`$ (80) It follows from Eq. (59) and $`B_0=z_0^2(0)`$ that $`\rho (0)=0`$ and hence $`\beta _0=0`$. Using the relation $$\mathrm{sn}(\alpha t|\nu )=\frac{1}{\sqrt{\nu }}\mathrm{sn}(\sqrt{\nu }\alpha t|\nu ^1),$$ (81) we can show that the two expressions in Eqs. (74), (77), and (80) are equivalent to each other. We used the lower expressions in Sec. III. The $`m=\pm 2`$ components $`\zeta _2(t)=\zeta _2(t)`$ have the same form as Eq. (69), in which $`A_0`$, $`\beta _0`$, and $`B_0`$ are replaced by $`A_2`$, $`\beta _2`$, and $`B_0`$. These constants are given by $$A_2=\{\begin{array}{c}\frac{1}{2}2\zeta _2^2(0)\frac{5q}{\stackrel{~}{c}_2},\hfill \\ \frac{\stackrel{~}{c}_2}{10q}\zeta _2^2(0)\left[12\zeta _2^2(0)\right],\hfill \end{array}$$ (82) $`\beta _2=0`$, and $`B_2=\zeta _2^2(0)`$. Finally, we consider the case of $`\chi _2=0`$ in the initial condition (25), i.e., all the components are real and positive. We can solve Eq. (61) also in this case to give $`B_0=z_0^2(0)`$ and hence $`\beta _0=0`$. The constants $`A_0`$, $`\nu `$, and $`\alpha `$ reduce to $`A_0`$ $`=`$ $`{\displaystyle \frac{1}{2\stackrel{~}{c}_2q}}\left\{q\left\{\stackrel{~}{c}_2[12\zeta _0^2(0)]+10q\right\}\right|q|\sqrt{(\stackrel{~}{c}_2+10q)^240\stackrel{~}{c}_2q\zeta _0^2(0)}\},`$ $`\nu `$ $`=`$ $`{\displaystyle \frac{q\left\{\stackrel{~}{c}_2[12\zeta _0^2(0)]+10q\right\}|q|\sqrt{(\stackrel{~}{c}_2+10q)^240\stackrel{~}{c}_2q\zeta _0^2(0)}}{q\left\{\stackrel{~}{c}_2[12\zeta _0^2(0)]+10q\right\}+|q|\sqrt{(\stackrel{~}{c}_2+10q)^240\stackrel{~}{c}_2q\zeta _0^2(0)}}},`$ (84) $`\alpha `$ $`=`$ $`\sqrt{{\displaystyle \frac{8\stackrel{~}{c}_2qA_0}{5\mathrm{}^2\nu }}},`$ (85) where we take the lower sign in Eq. (A). Since $`\left\{\stackrel{~}{c}_2[12\zeta _0^2(0)]+10q\right\}^2<(\stackrel{~}{c}_2+10q)^240\stackrel{~}{c}_2q\zeta _0^2(0)`$, we find $`\stackrel{~}{c}_2A>0`$ and $`\nu <0`$.
warning/0506/hep-th0506008.html
ar5iv
text
# 1 Introduction ## 1 Introduction Noncommutative geometry has suggested the introduction of *fuzzy* spaces, namely the approximation of the abelian algebra of functions on an ordinary space, with a finite rank matrix algebra, which preserves the symmetries of the original space, at the price of noncommutativity. The first formalization of the idea of a fuzzy space was introduced by Madore , for the sphere. In his approach, a fuzzy sphere is a sequence of nonabelian algebras, generated (loosely speaking) by three “noncommutative coordinates” which satisfy $`x_ix_i=1`$ and $`[x_i,x_j]=\kappa \epsilon _{ijk}x_k`$, with $`\kappa `$ some noncommutativity parameter. The connection with the finite dimensional representations of $`SU(2)`$ is immediate, and with $`\kappa `$ proportional to the inverse of the rank of the representation, one can think of recovering functions on the sphere in some limit. This limit has been made rigorous by Rieffel at the level of topological “quantum spaces”, but here we are more interested in the approximations that the fuzzy matrix geometry can give to the space of functions (and hence of fields) defined on some space. This is efficiently done defining “fuzzy spherical harmonics” , a basis for the matrix algebra, which in the limit converges to the ordinary spherical harmonics. Field theory on the fuzzy sphere is a very active field of research. A partial list of references is . After the fuzzy sphere, similar approximations for higher dimensional spheres , for the torus, based on the noncommutative torus algebra , and for projective spaces , have also been studied. In previous work , we proposed a fuzzy approximations for the algebra of functions on a disc (see also ). It is the first example of a fuzzy approximation for a space that classically has a boundary. The space is defined through a suitable truncation of the noncommutative plane. In this paper we examine the spectral properties of the fuzzy Laplacian, finding its eigenvalues and eigenvectors, and comparing them with the exact case and give a direct way to associate, to functions on the disc, their fuzzy counterpart, i.e. to approximate the abelian algebra of functions by a sequence of finite rank matrices. We introduce a basis for the algebra of functions on the fuzzy disc in terms of symbols of the eigenmatrices of the fuzzy approximation to the Laplacian. These symbols are seen to converge, as the rank of the matrices increases, towards a set of functions on the plane which coincide with the ordinary Bessel functions of first kind (eigenfunctions of the ordinary Laplacian with Dirichlet boundary conditions) for points on a plane inside a disc, and to zero for points outside this disc. These are given the name of fuzzy Bessel functions. The paper is organised as follows. In the next section we review the fuzzy sphere from a perspective which will render easy the generalization to the disc case. The main tool we use is the Weyl-Wigner formalism, and the concept of generalized coherent sates, which we briefly review in appendices A and B respectively. In section 3 we review the fuzzy disc. A Weyl-Wigner isomorphism is introduced in terms of functions on the plane, where noncommutativity is represented by a parameter $`\theta `$. In this formalization, there is no natural concept of a sequence of finite dimensional Hilbert spaces, or finite rank matrix algebras, therefore a truncation in the algebra of operators is introduced, with respect to a specific basis in the Hilbert space. Constraining the dimension $`N`$ of the truncation to the noncommutativity parameter, $`N\theta =R^2`$, one obtains then a sequence of finite rank matrices, converging towards an abelian algebra of operators, that approximates functions whose support is concentrated on a disc of radius $`R`$. On this sequence of states, fuzzy derivatives and a fuzzy Laplacian are defined and the spectral properties of the latter are studied. Then in section 4 we study eigenvalues and eigenvectors and show that they (rather quickly) converge to the ordinary Bessel functions. ## 2 The fuzzy sphere in the Weyl-Wigner formalism In this section we present the fuzzy sphere in the general context of the Weyl-Wigner formalism . We first set up an isomorphism between a space of operators and a space of functions on a sphere. Since a sphere is the coadjoint orbit of the group $`SU\left(2\right)`$, the basic tool to introduce this map is a system of coherent states, specialising the general arguments of appendix B. To define a system of coherent states consider the unitary irreducible representations for the group $`SU\left(2\right)`$. On each finite dimensional Hilbert space $`\text{}^N`$, with $`N=2L+1`$ -here $`\left(L=0,1/2,1,3/2\mathrm{}\right)`$-, a basis is given by vectors that, in ket notation, are represented as $`|L,M`$ -with $`M=(L,L+1,\mathrm{},L1,L)`$-. One has: $$uSU\left(2\right)\stackrel{\widehat{R}^{\left(L\right)}}{}\left(\text{}^N\right).$$ (2.1) The matrix elements of this representation are given by: $$L,M|\widehat{R}^{\left(L\right)}\left(u\right)|L,M^{}=D_{MM^{}}^L\left(u\right).$$ (2.2) These are called Wigner functions . The second step is to fix a fiducial state. One can choose the highest weight in the representation: $`|\psi _0=|L,L`$. If the group manifold is parametrised by Euler angles, then $`u`$ represents a point whose “coordinates” range through $`\alpha [0,4\pi )`$,$`\beta [0,\pi )`$, $`\gamma [0,2\pi )`$. Fixed the fiducial vector, its stability subgroup $`H_{\psi _0}`$ by the $`\widehat{R}^{\left(L\right)}`$ representation is made by elements for which $`\beta =0`$ (this condition is seen to be valid independently of $`N`$). Two elements $`u`$ and $`u^{}`$ are equivalent if $`u^{}u^{}H_{\psi _0}`$. It is possible to prove that: $$SU\left(2\right)/H_{\psi _0}S^2$$ (2.3) identifying $`\theta =\beta `$ and $`\phi =\alpha mod\mathrm{\hspace{0.17em}2}\pi `$. Chosen a representative element $`\stackrel{~}{u}`$ in each equivalence class of the quotient, the set of coherent states is defined as: $$|\theta ,\phi ,N=\widehat{R}^{\left(L\right)}\left(\stackrel{~}{u}\right)|L,L.$$ (2.4) The left hand side ket now explicitly depends on $`N`$, the dimension of the space on which the representation takes place. Projecting on the basis elements, one has: $`L,M|\theta ,\phi ,N`$ $`=`$ $`D_{ML}^L\left(\stackrel{~}{u}\right),`$ (2.5) $`|\theta ,\phi ,N`$ $`=`$ $`{\displaystyle \underset{M=L}{\overset{L}{}}}\left(\frac{\left(2L\right)!}{\left(L+M\right)!\left(LM\right)!}\right)^{1/2}\left(\mathrm{cos}\theta /2\right)^{L+M}\left(\mathrm{sin}\theta /2\right)^{LM}e^{i\phi M}|L,M.`$ This set of states is nonorthogonal, and overcomplete ($`d\mathrm{\Omega }=d\phi \mathrm{sin}\theta d\theta `$): $`\theta ^{},\phi ^{},N|\theta ,\phi ,N`$ $`=`$ $`e^{iL\left(\phi ^{}\phi \right)}\left[e^{i\left(\phi ^{}\phi \right)}\mathrm{cos}\theta /2\mathrm{cos}\theta ^{}/2+\mathrm{sin}\theta /2\mathrm{sin}\theta ^{}/2\right]^{2L},`$ $`11`$ $`=`$ $`{\displaystyle \frac{2L+1}{4\pi }}{\displaystyle _{S^2}}𝑑\mathrm{\Omega }|\theta ,\phi ,N\theta ,\phi ,N|.`$ (2.6) Using this set of vectors it is possible to define a map, from the space of operators on a finite dimensional Hilbert space to the space of functions on the sphere $`S^2`$: $`\widehat{A}^{\left(N\right)}\left(\text{}^N\right)𝕄_N`$ $``$ $`A^{\left(N\right)}\left(S^2\right),`$ $`A^{\left(N\right)}(\theta ,\phi )`$ $`=`$ $`\theta ,\phi ,N|\widehat{A}^{\left(N\right)}|\theta ,\phi ,N.`$ (2.7) This is a way to map every finite rank matrix into a function on a sphere, called the Berezin symbol of the given matrix . Among these operators, there are $`\widehat{Y}_{JM}^{\left(N\right)}`$ whose symbols are the spherical harmonics, up to order $`2L`$ (here $`J=0,1,\mathrm{},2L`$ and $`M=J,\mathrm{},+J`$): $$\theta ,\phi ,N|\widehat{Y}_{JM}^{\left(N\right)}|\theta ,\phi ,N=Y_{JM}(\theta ,\phi ),$$ (2.8) these operators are called *fuzzy harmonics*. The origin of this name must be traced back to the definition, on each finite rank matrix algebra $`𝕄_N`$, of an operator, in terms of the generators $`\widehat{L}_a^{\left(N\right)}`$: $$[\widehat{L}_a^{\left(N\right)},\widehat{L}_b^{\left(N\right)}]=iϵ_{abc}\widehat{L}_c^{\left(N\right)},$$ (2.9) representing the Lie algebra of the group $`SU\left(2\right)`$ on the space $`\text{}^N`$: $`^2`$ $`:`$ $`𝕄_N𝕄_N,`$ $`^2\widehat{A}^{\left(N\right)}`$ $`=`$ $`[\widehat{L}_s^{\left(N\right)},[\widehat{L}_s^{\left(N\right)},\widehat{A}^{\left(N\right)}]].`$ (2.10) This operator is called the *fuzzy Laplacian*. Its spectrum is given by eigenvalues $`_j=j\left(j+1\right)`$, where $`j=0,\mathrm{},2L`$, and every eigenvalue has a multiplicity of $`2j+1`$. The spectrum of the fuzzy Laplacian thus coincides, up to order $`2L`$, with the one of its continuum counterpart acting on the space of functions on a sphere. The cut-off of this spectrum is of course related to the dimension of the rank of the matrix algebra under analysis. The fuzzy harmonics are the eigenstates of this operator. Fuzzy harmonics are a basis in each space of matrices $`𝕄_N`$. They are orthogonal but non normalized and are proportional to the polarization operators $`\widehat{T}_{JM}^{\left(N\right)}`$ defined in terms of the Clebsh-Gordan coefficients for $`SU(2)`$ : $`[\widehat{L}_\mu ^{\left(N\right)},\widehat{T}_{JM}^{\left(N\right)}]`$ $`=`$ $`\sqrt{L\left(L+1\right)}C_{LM1\mu }^{LM+\mu }\widehat{T}_{LM+\mu }^{\left(N\right)},`$ $`Tr\left[\widehat{T}_{JM}^{\left(N\right)}\widehat{T}_{J^{}M^{}}^{\left(N\right)}\right]`$ $`=`$ $`\delta _{JJ^{}}\delta _{MM^{}},`$ $`\widehat{T}_{JM}^{\left(N\right)}`$ $`=`$ $`\left(1\right)^M\widehat{T}_{JM}^{\left(N\right)}.`$ (2.11) the proportionality being: $$\widehat{Y}_{JM}^{\left(N\right)}=\frac{J!}{\left(2lJ\right)^J}\left[\frac{\left(2l+J+1\right)!}{4\pi \left(2lJ\right)!}\right]^{1/2}\widehat{T}_{JM}^{\left(N\right)},$$ (2.12) from which: $$\mathrm{Tr}\left[\widehat{Y}_{JM}^{\left(N\right)}\widehat{Y}_{J^{}M^{}}^{\left(N\right)}\right]=\left(\frac{J!}{\left(2lJ\right)^J}\right)^2\left(\frac{\left(2l+J+1\right)!}{4\pi \left(2lJ\right)!}\right)\delta _{JJ^{}}\delta _{MM^{}}.$$ (2.13) Using this basis, an element $`\widehat{F}^{\left(N\right)}`$ belonging to $`𝕄_N`$ can be expanded as: $$\widehat{F}^{\left(N\right)}=\underset{J=0}{\overset{2L}{}}\underset{M=J}{\overset{J}{}}F_{JM}^{\left(N\right)}\widehat{Y}_{JM}^{\left(N\right)},$$ (2.14) with coefficients $$F_{JM}^{\left(N\right)}=\frac{\mathrm{Tr}\left[\widehat{Y}_{JM}^{\left(N\right)}\widehat{F}^{\left(N\right)}\right]}{\mathrm{Tr}\widehat{Y}_{JM}^{\left(N\right)}\widehat{Y}_{JM}^{\left(N\right)}}.$$ (2.15) A Weyl-Wigner map can be defined simply mapping spherical harmonics into fuzzy harmonics: $$\widehat{Y}_{JM}^{\left(N\right)}Y_{JM}(\theta ,\phi ).$$ (2.16) This map clearly depends on the dimension $`N`$ of the space on which fuzzy harmonics are realized. It can be linearly extended by: $$\widehat{F}^{\left(N\right)}=\underset{J=0}{\overset{2L}{}}\underset{M=J}{\overset{J}{}}F_{JM}^{\left(N\right)}\widehat{Y}_{JM}^{\left(N\right)}F^{\left(N\right)}(\theta ,\varphi )=\underset{J=0}{\overset{2L}{}}\underset{M=J}{\overset{+J}{}}F_{JM}^{\left(N\right)}Y_{JM}(\theta ,\varphi ).$$ (2.17) This is a Weyl-Wigner isomorphism and it can be used to define a fuzzy sphere. Given a function on a sphere, if it is square integrable with respect to the standard measure $`d\mathrm{\Omega }`$, then it can be expanded in the basis of spherical harmonics: $$f(\theta ,\phi )=\underset{J=0}{\overset{\mathrm{}}{}}\underset{M=J}{\overset{J}{}}f_{JM}Y_{JM}(\theta ,\phi ).$$ (2.18) Now consider the set of “truncated” functions: $$f^{\left(N\right)}(\theta ,\phi )=\underset{J=0}{\overset{2L}{}}\underset{M=J}{\overset{J}{}}f_{JM}Y_{JM}(\theta ,\phi ).$$ (2.19) This is a vector space, but it is no more an algebra, with the standard definition of sum and pointwise product of two functions, as the product of two spherical harmonics of order say $`2L`$ has spherical components of order larger than $`2L`$. We have indeed: $$\left(Y_{J^{}M^{}}Y_{J^{\prime \prime }M^{\prime \prime }}\right)(\theta ,\phi )=\underset{J=\left|J^{}J^{\prime \prime }\right|}{\overset{J=\left|J^{}+J^{\prime \prime }\right|}{}}\underset{M=J}{\overset{J}{}}\sqrt{\frac{\left(2J^{}+1\right)\left(2J^{\prime \prime }+1\right)}{4\pi \left(2J+1\right)}}C_{J^{}0J^{\prime \prime }0}^{J0}C_{J^{}M^{}J^{\prime \prime }M^{\prime \prime }}^{JM}Y_{JM}(\theta ,\phi )$$ (2.20) in terms of Clebsh-Gordan coefficients for $`SU\left(2\right)`$. If these truncated functions are mapped, via the Weyl-Wigner procedure, into matrices: $$f^{\left(N\right)}(\theta ,\phi )\widehat{f}^{\left(N\right)}=\underset{J=0}{\overset{2L}{}}\underset{M=J}{\overset{+J}{}}f_{JM}\widehat{Y}_{JM}^{\left(N\right)},$$ (2.21) then this set of truncated functions is given the vector space structure of $`𝕄_N`$. But $`𝕄_N`$ is even a *nonabelian* algebra. Invertibility of this association (2.21) enables to define, in the set of the truncated functions on the sphere, a non abelian product, isomorphic to that of matrices: $$\left(f^{\left(N\right)}g^{\left(N\right)}\right)(\theta ,\phi )=\underset{J=0}{\overset{2L}{}}\underset{M=J}{\overset{+J}{}}\frac{Tr\left[\widehat{f}^{\left(N\right)}\widehat{g}^{\left(N\right)}\widehat{Y}_{JM}^{\left(N\right)}\right]Y_{JM}(\theta ,\phi )}{\mathrm{Tr}\widehat{Y}_{JM}^{\left(N\right)}\widehat{Y}_{JM}^{\left(N\right)}}.$$ (2.22) The Weyl-Wigner map (2.17) has been used to make each set of truncated functions a non abelian algebra $`𝒜^{\left(N\right)}(S^2,)`$, isomorphic to $`𝕄_N`$. The product of two fuzzy harmonics can be obtained in term of the 6j-symbols from the product of two polarization operators $$\widehat{T}_{J^{}M^{}}^{\left(N\right)}\widehat{T}_{J^{\prime \prime }M^{\prime \prime }}^{\left(N\right)}=\underset{J}{}\left(1\right)^{2L+J}\sqrt{\left(2J^{}+1\right)\left(2J^{\prime \prime }+1\right)}\left\{\begin{array}{ccc}J^{}& J^{\prime \prime }& J\\ L& L& L\end{array}\right\}C_{J^{}M^{}J^{\prime \prime }M^{\prime \prime }}^{JM}\widehat{T}_{JM}^{\left(N\right)}.$$ (2.23) These algebras can be seen as formally generated by matrices which are the images of the norm $`1`$ vectors in $`\text{}^3`$, i.e. points on a sphere. They are mapped into multiples of the generators $`\widehat{L}_a^{\left(N\right)}`$ of the Lie algebra: $$\frac{x_a}{\stackrel{}{x}}\widehat{x}_a^{\left(N\right)}[\widehat{x}_a^{\left(N\right)},\widehat{x}_b^{\left(N\right)}]=\frac{2i\epsilon _{abc}}{\sqrt{N^21}}\widehat{x}_c^{\left(N\right)}.$$ (2.24) The commutation rules satisfied by generators of the algebras in the sequence $`𝒜^{\left(N\right)}(S^2,)`$ make it intuitively clear that the limit for $`N\mathrm{}`$ of this sequence is an abelian algebra. This is the reason why this sequence is called *fuzzy sphere*. ## 3 The fuzzy disc In the previous section, the fuzzy sphere has been introduced via the Weyl-Wigner formalism, which makes use of the fact that the 2-dimensional sphere is the coadjoint orbit of the *compact* group $`SU\left(2\right)`$. Compactness of the group brought to a natural definition of a sequence of finite dimensional Hilbert spaces, and a natural identification of the dimension of these spaces as a cut-off index for a suitable expansion of a generic function on the sphere. Properties of the sphere as an orbit of that group made it possible to formalize every point of the sphere as a coherent state, defined on each of those finite dimensional spaces carrying the UIRR’s. So the map between operators (finite rank matrices) and functions on the sphere has been naturally introduced via the Berezin procedure, and has been refined stressing the role of the fuzzy Laplacian operator in the definition of fuzzy harmonics as a basis in the set $`𝕄_N`$. Moreover, compactness of the group played a fundamental role in the rigorous analysis of the convergence performed by Rieffel . To approach the problem of defining a fuzzy approximation to the disc, it would be natural then to analyse the possibility that the disc were a coadjoint orbit for a Lie group. If this were the case, it would be possible to introduce a system of coherent states labelled by its points, and some sort of Berezin map . However this approach meets some troubles. The group from which it is possible to define a system of coherent states in correspondence with points of a disc is $`SU(1,1)`$, but it is non compact, so its UIRR’s are not realized on finite dimensional Hilbert spaces. Hence, there is no intrinsic concept of a cut-off index, the would be dimension of the fuzzification, in this context. In the problem has been circumvented using coherent states for the Heisenberg-Weyl group and the standard isomorphism in terms of Berezin symbols between operators and functions on the plane. Then a sequence of projections converging to the characteristic function of the disc has been defined. The projections identify a sequence of finite rank matrix algebras which can be endowed with an additional structure, a fuzzy Laplacian, identifying the underlying geometry. In the commutative limit $`N\theta =R^2`$ this geometry is seen to converge to a disc of radius $`R`$. It is well known that it is possible to define a basis for the space of functions on the disc in terms of a suitable set of Bessel functions. Since the Weyl-Wigner map for functions on the sphere has been introduced (2.17) mapping spherical harmonics into fuzzy harmonics, we pose the question whether or not it is possible to define a system of *fuzzy Bessel* functions in terms of finite rank matrices, and then map again continuum Bessel functions into them. Since Bessel functions are defined on the plane, in connection with boundary values problems for Laplacian operators, the noncommutative plane which we are going to revue, together with fuzzy derivations and a fuzzy Laplacian will turn out to be a natural setting to perform the analysis. The final answer to the problem will be given studying the spectral resolution of the fuzzy Laplacian. ### 3.1 The noncommutative plane as a matrix algebra The noncommutative plane can be defined using a Weyl-Wigner map, following again the general procedure of Berezin: so the first step will be the definition of a set of generalised coherent states for the Heisenberg-Weyl group, since they are labelled by points of a plane. In this space it will be assumed a system of coordinates of the kind $`(x,y)`$ for a point of $`\text{}^2`$; $`\theta `$ will be a parameter representing an explicit noncommutativity. With this notation, Heisenberg-Weyl group is a manifold $`\text{}^3`$, whose points are represented by a triple $`(x,y,\lambda )`$, with the composition rule: $$(x,y,\lambda )(x^{},y^{},\lambda ^{})=(x+x^{},y+y^{},\lambda +\lambda ^{}\frac{1}{\theta }\left(xy^{}x^{}y\right)).$$ (3.25) The identity of this group is given by: $$id_W=(0,0,0),$$ (3.26) and the inverse of a generic element is: $$(x,y,\lambda )^1=(x,y,\lambda ).$$ (3.27) Considering the complex plane via $`z=x+iy`$ and $`\overline{z}=xiy`$ the group law becomes: $$(z,\lambda )(z^{},\lambda ^{})=(z+z^{},\lambda +\lambda ^{}+\frac{i}{2\theta }\left(\overline{z}z^{}\overline{z}^{}z\right)).$$ (3.28) The Hilbert space on which representing this group is again the Fock space $``$ of finite norm complex analytical functions in the $`w`$ variable where the norm is obtained by the scalar product: $$f|g\frac{d^2w}{\pi \theta }e^{\overline{w}w/\theta }\overline{f}\left(w\right)g\left(w\right),$$ (3.29) and an orthonormal basis is given by: $$\psi _n\left(w\right)=\frac{w^n}{\sqrt{\theta ^nn!}}.$$ (3.30) The unitary representation of the Heisenberg-Weyl group is given by operators $`\widehat{T}(z,\lambda )Op\left(\right)`$: $$\left(\widehat{T}f\right)\left(w\right)=e^{i\lambda }e^{\overline{z}z/2\theta }e^{zw/\theta }f\left(w\overline{z}\right).$$ (3.31) Chosen the first basis element $`\psi _0\left(w\right)`$ as the fiducial state, the procedure already outlined gives a coherent state for each point of the complex plane. In the realization of the Fock space as a space of complex analytical functions, a coherent state is then: $$|z\psi _z\left(w\right)=e^{\overline{z}z/2\theta }e^{zw/\theta }$$ (3.32) and, given $`\psi _n`$ an element of the basis already considered: $$|z=\underset{n=0}{\overset{\mathrm{}}{}}e^{\overline{z}z/2\theta }\frac{z^n}{\sqrt{n!\theta ^n}}|\psi _n.$$ (3.33) Such coherent states are non orthogonal, and overcomplete: $`z|z^{}`$ $`=`$ $`e^{\left(\left|z\right|^2+\left|z^{}\right|^22\overline{z}z^{}\right)/\theta },`$ $`11`$ $`=`$ $`{\displaystyle \frac{d^2z}{\pi \theta }|zz|}.`$ (3.34) On this Hilbert space $``$, it is possible to introduce a pair of creation-annihilation operators: $`\left(\widehat{a}f\right)\left(w\right)`$ $`=`$ $`\theta {\displaystyle \frac{df}{dw}}`$ $`\left(\widehat{a}^{}f\right)\left(w\right)`$ $`=`$ $`wf\left(w\right),`$ (3.35) such that: $$[\widehat{a},\widehat{a}^{}]=\theta \mathrm{\hspace{0.17em}1}1,$$ (3.36) and: $$\widehat{a}|z=z|z,$$ (3.37) so that we can recognize the usual coherent states encountered in quantum mechanics textbooks. In the chosen orthonormal basis, one has: $`\widehat{a}|\psi _n`$ $`=`$ $`\sqrt{n\theta }|\psi _{n1}`$ $`\widehat{a}^{}|\psi _n`$ $`=`$ $`\sqrt{\left(n+1\right)\theta }|\psi _{n+1},`$ (3.38) and therefore one recognizes the basis as the basis of eigenvectors of the number operator $`\widehat{N}=\widehat{a}^{}\widehat{a}`$. Note however the unconventional presence $`\theta `$ in the commutation relation (3.36). These relations can be extended. A Berezin symbol can be associated to an operator in the Fock space: $$f(\overline{z},z)=z|\widehat{f}|z.$$ (3.39) This can be seen as a Wigner map. It can be inverted: $$\widehat{f}=\frac{d^2\xi }{\pi \theta }\frac{d^2z}{\pi \theta }f(z,\overline{z})e^{\left(\overline{z}\xi \overline{\xi }z\right)/\theta }e^{\xi \widehat{a}^{}/\theta }e^{\overline{\xi }\widehat{a}/\theta }.$$ (3.40) This quantization map for functions on a plane can be given an interesting form. To start with, we can restrict to functions which can be written as Taylor series in $`\overline{z},z`$: $$f(\overline{z},z)=\underset{m,n=0}{\overset{\mathrm{}}{}}f_{mn}^{Tay}\overline{z}^mz^n.$$ (3.41) An easy calculation shows that this $`f`$ is the symbol of the operator: $$\widehat{f}=\underset{m,n=0}{\overset{\mathrm{}}{}}f_{mn}^{Tay}\widehat{a}^m\widehat{a}^n.$$ (3.42) More generally we can consider operators written in a density matrix notation: $$\widehat{f}=\underset{m,n=0}{\overset{\mathrm{}}{}}f_{mn}|\psi _m\psi _n|.$$ (3.43) The Berezin symbol of this operator is the function: $$f(\overline{z},z)=e^{\left|z\right|^2/\theta }\underset{m,n=0}{\overset{\mathrm{}}{}}f_{mn}\frac{\overline{z}^mz^n}{\sqrt{m!n!\theta ^{m+n}}},$$ (3.44) where the relation between the Taylor coefficients $`f_{mn}^{Tay}`$ and the $`f_{mn}`$ is $$f_{lk}=\underset{q=0}{\overset{\mathrm{min}(l,k)}{}}f_{lqkq}^{Tay}\frac{\sqrt{k!l!\theta ^{l+k}}}{q!\theta ^q},$$ (3.45) while the inverse relation is given by: $$f_{mn}^{Tay}=\underset{p=0}{\overset{\mathrm{min}(m,n)}{}}\frac{\left(1\right)^p}{p!\sqrt{\left(mp\right)!\left(np\right)!\theta ^{m+n}}}f_{mpnp}.$$ (3.46) Equation (3.42) shows that the quantization of a monomial in the variables $`z,\overline{z}`$ is an operator in $`\widehat{a},\widehat{a}^{}`$, formally a monomial in these two noncommuting variables, with all terms in $`\widehat{a}^{}`$ acting at the left side with respect to terms in $`\widehat{a}`$. This ordering is related to the presence, in the quantization map (3.40), of a specific term that, in the standard terminology used for the Weyl-Wigner formalism, is called weight . In fact, restoring real variables (3.40) can be written as: $$\widehat{f}=\frac{dadb}{2\pi \theta }\frac{dxdy}{2\pi \theta }f(x,y)e^{\frac{i\left(bxay\right)}{\theta }}e^{\frac{u\widehat{a}^{}}{\theta }}e^{\frac{\overline{u}\widehat{a}}{\theta }}=\frac{dadb}{2\pi \theta }\stackrel{~}{f}(b,a)e^{\frac{\left(a^2+b^2\right)}{8\theta }}e^{\frac{\left(u\widehat{a}^{}\overline{u}\widehat{a}\right)}{\theta }}.$$ (3.47) with $`u=\left(a+ib\right)/2`$. The function $`\stackrel{~}{f}`$ is the *symplectic* Fourier transform of the function $`f`$. Compared (3.47) with the standard Weyl map , the difference is the factor $`e^{\left(a^2+b^2\right)/8\theta }`$. The invertibility of the Weyl map (on a suitable domain of functions on the plane) enables to define a noncommutative product in the space of functions, known as Voros product , a variant of the more popular Grönewold-Moyal product : $$\left(fg\right)(\overline{z},z)=z|\widehat{f}\widehat{g}|z.$$ (3.48) It is a non local product: $$\left(fg\right)(\overline{z},z)=e^{\overline{z}z/\theta }\frac{d^2\xi }{\pi \theta }f(\overline{z},\xi )g(\overline{\xi },z)e^{\overline{\xi }\xi /\theta }e^{\overline{\xi }z/\theta }e^{\overline{z}\xi /\theta }.$$ (3.49) Its asymptotic expansion, on a suitable domain, acquires the form: $$\left(fg\right)(\overline{z},z)=fe^{\theta \stackrel{}{}_{\overline{z}}\stackrel{}{}_z}g,$$ (3.50) and makes it clear that it is a deformation, in $`\theta `$, of the pointwise commutative product. Since it is the translation, in the space of functions, of the product in the space of operators, if symbols are expressed in the form (3.44), then the product acquires a matrix form: $$\left(fg\right)_{mn}=\underset{k=0}{\overset{\mathrm{}}{}}f_{mk}g_{kn}.$$ (3.51) The space of functions on the plane, with the standard definition of sum, and the product given by the Voros product (3.48), is a nonabelian algebra, a noncommutative plane. This algebra $`𝒜_\theta =(\left(\text{}^2\right),)`$ is isomorphic to an algebra of operators, or, as equation (3.51) suggests, to an algebra of infinite dimensional matrices. A more detailed mathematical analysis of the Moyal algebra, i.e. of the quantum plane, is given in and references therein. ### 3.2 A sequence of non abelian algebras A fuzzy space has been presented as a sequence of finite rank matrix algebras converging to an algebra of functions. The meaning of this convergence is formalised as compact quantum metric spaces. In the case of the fuzzy sphere the rank of the matrices involved is the dimension of the Hilbert space on which UIRR’s of the group $`SU\left(2\right)`$ are realised. In the approach sketched in the last section there is no natural definition of a set of finite dimensional matrix algebras. This is a consequence of the fact that the noncommutative plane has been realised via a Berezin quantization based on coherent states originated by a group, the Heisenberg-Weyl, which is noncompact. In this context the strategy to obtain finite dimensional matrix algebras is different. $`𝒜_\theta `$ can be considered, once a basis in the Hilbert space has been chosen, as a matrix algebra made up by formally infinite dimensional matrices. One can define a set of finite dimensional matrix algebras simply truncating $`𝒜_\theta `$. The notion of truncation is formalised via the introduction of a set of projectors: $$\widehat{P}_\theta ^{(N)}=\underset{n=0}{\overset{N}{}}|\psi _n\psi _n|$$ (3.52) in the space of operators. Their symbols are projectors in the algebra $`𝒜_\theta `$ of the noncommutative plane, in the sense that they are idempotent functions of order $`2`$ with respect to the Voros product (here $`z=re^{i\phi }`$): $`P_\theta ^{\left(N\right)}(r,\phi )`$ $`=`$ $`{\displaystyle \underset{n=0}{\overset{N}{}}}z|\psi _n\psi _n|z=e^{r^2/\theta }{\displaystyle \underset{n=0}{\overset{N}{}}}{\displaystyle \frac{r^{2n}}{n!\theta ^n}}`$ $`P_\theta ^{\left(N\right)}P_\theta ^{\left(N\right)}`$ $`=`$ $`P_\theta ^{\left(N\right)}.`$ (3.53) This finite sum can be performed yielding a rotationally symmetric function: $$P_\theta ^{\left(N\right)}(r,\phi )=\frac{\mathrm{\Gamma }(N+1,r^2/\theta )}{\mathrm{\Gamma }\left(N+1\right)}.$$ (3.54) in terms of the ratio of an incomplete gamma function by a gamma function . If $`\theta `$ is kept fixed, and nonzero, in the limit for $`N\mathrm{}`$ the symbol $`P_\theta ^{\left(N\right)}(r,\phi )`$ converges, pointwise, to the constant function $`P_\theta ^{\left(N\right)}(r,\phi )=1`$, which can be formally considered as the symbol of the identity operator: in this limit one recovers the whole noncommutative plane. This situation changes if the limit for $`N\mathrm{}`$ is performed keeping the product $`N\theta `$ equals to a constant, say $`R^2`$. In a pointwise convergence, chosen $`R^2=1`$: $$P_\theta ^{\left(N\right)}\left[\begin{array}{cc}1& r<1\\ 1/2& r=1\\ 0& r>1\end{array}\right]=Id\left(r\right).$$ (3.55) This sequence of projectors converges to a step function in the radial coordinate $`r`$, the characteristic function of a disc on the plane. Thus a sequence of subalgebras $`𝒜_\theta ^{\left(N\right)}`$ can be defined by: $$𝒜_\theta ^{\left(N\right)}=P_\theta ^{\left(N\right)}𝒜_\theta P_\theta ^{\left(N\right)}.$$ (3.56) As it has been said, the full algebra $`𝒜_\theta `$ is isomorphic to an algebra of operators. What the previous relation says is that $`𝒜_\theta ^{\left(N\right)}`$ is isomorphic to $`𝕄_{N+1}`$, the algebra of $`\left(N+1\right)`$ rank matrices: the important thing is that this isomorphism (this truncation) is obtained via a specific choice of a basis in the Fock space $``$ on which coherent states for the Heisenberg-Weyl group are realised. The effect of this projection on a generic function is: $$\mathrm{\Pi }_\theta ^{\left(N\right)}\left(f\right)=f_\theta ^{\left(N\right)}=P_\theta ^{\left(N\right)}fP_\theta ^{\left(N\right)}=e^{\left|z\right|^2/\theta }\underset{m,n=0}{\overset{N}{}}f_{mn}\frac{\overline{z}^mz^n}{\sqrt{m!n!\theta ^{m+n}}}.$$ (3.57) On every subalgebra $`𝒜_\theta ^{\left(N\right)}`$, the symbol $`P_\theta ^{\left(N\right)}(r,\phi )`$ is then the identity, because it is the symbol of the projector $`\widehat{P}^{\left(N\right)}=_{n=0}^N|\psi _n\psi _n|`$, which is the identity operator in $`𝒜_\theta ^{\left(N\right)}`$, or, equivalently, the identity matrix in every $`𝕄_{N+1}`$. Note that the rotation group on the plane, $`SO\left(2\right)`$, acts in a natural way on these subalgebras. Its generator is the truncated number operator $`\widehat{N}^{(N)}=_{n=0}^Nn\theta |\psi _n\psi _n|`$. Cutting at a finite $`N`$ the expansion provides an infrared cutoff. This cutoff is “fuzzy” in the sense that functions in the subalgebra are still defined outside the disc of radius $`R`$, but are exponentially damped. In general functions are close to their projected version $`f_\theta ^{\left(N\right)}`$ if they are mostly supported on a disc of radius $`R=\sqrt{N\theta }`$, otherwise they are exponentially cut, provided they do not present oscillations of too small wavelength (compared to $`\theta `$). In this case the projected function becomes very large on the boundary of the disc. More details and examples are in . ### 3.3 Fuzzy derivatives and fuzzy Laplacian So far the Weyl-Wigner formalism, and the projection procedure, have provided a way to associate to functions on the plane a sequence of finite dimension $`\left(N+1\right)\times \left(N+1\right)`$ matrices. An appropriate choice of $`\theta `$, the noncommutativity parameter introduced by the quantization map, and $`N`$, showed that it is possible to obtain a good approximation of a certain class of functions supported on a disc. The next step is the analysis of the geometry these algebras can formalize. To pursue this task, we need derivations and a Laplacian. The starting point to define the matrix equivalent of the derivations is: $`_zf`$ $`=`$ $`{\displaystyle \frac{1}{\theta }}z|[\widehat{f},\widehat{a}^{}]|z,`$ $`_{\overline{z}}f`$ $`=`$ $`{\displaystyle \frac{1}{\theta }}z|[\widehat{a},\widehat{f}]|z.`$ (3.58) This relation is exact in the full algebra $`𝒜_\theta `$. Given an operator $`\widehat{f}`$, the derivatives of the symbol $`f(\overline{z},z)`$ are related to the symbol of the commutator of $`\widehat{f}`$ with the creation and annihilation operators. A fuzzified version, namely a truncated version, of these operations, can be defined as<sup>1</sup><sup>1</sup>1This a slightly modified version of the derivations in : $`_zf_\theta ^{\left(N\right)}`$ $``$ $`{\displaystyle \frac{1}{\theta }}z|\widehat{P}_\theta ^{\left(N\right)}[\widehat{P}_\theta ^{\left(N\right)}\widehat{f}\widehat{P}_\theta ^{\left(N\right)},\widehat{a}^{}]\widehat{P}_\theta ^{\left(N\right)}|z`$ $`_{\overline{z}}f_\theta ^{\left(N\right)}`$ $``$ $`{\displaystyle \frac{1}{\theta }}z|\widehat{P}_\theta ^{\left(N\right)}[\widehat{P}_\theta ^{\left(N\right)}\widehat{f}\widehat{P}_\theta ^{\left(N\right)},\widehat{a}]\widehat{P}_\theta ^{\left(N\right)}|z.`$ (3.59) It is important to note that this is really a derivation on each $`𝒜_\theta ^{\left(N\right)}`$, that is a linear operation from $`𝒜_\theta ^{\left(N\right)}`$ to itself, satisfying the Leibnitz rule. The way this is achieved in (3.59) goes through the following steps. We first associate to $`f_\theta ^{\left(N\right)}`$ a finite rank matrix $`\widehat{f}_\theta ^{\left(N\right)}`$ which is indeed regarded as embedded in an infinite dimensional one, with all zeroes but $`(N+1)\times (N+1)`$ elements in the upper left corner: $`\widehat{f}_\theta ^{\left(N\right)}=\widehat{P}_\theta ^{\left(N\right)}\widehat{f}\widehat{P}_\theta ^{\left(N\right)}`$. We then perform the commutator with $`\widehat{a},\widehat{a}^{}`$ realised as infinite dimensional matrices. This operation yields an infinite dimensional matrix with all zeroes but $`(N+2)\times (N+2)`$ elements in the upper left corner, so that we have to project back to $`𝕄_{N+1}`$ to obtain a properly defined derivation. Finally, we consider the symbol. As an example let us calculate the derivatives of the fuzzified coordinate functions $`z`$ and $`\overline{z}`$. From the definition, $`z`$, $`\overline{z}`$ are respectively the symbol of the annihilation and creation operators, which can be written as: $`\widehat{a}`$ $`=`$ $`{\displaystyle \underset{s=0}{\overset{\mathrm{}}{}}}\sqrt{\left(s+1\right)\theta }|\psi _s\psi _{s+1}|`$ $`\widehat{a}^{}`$ $`=`$ $`{\displaystyle \underset{k=0}{\overset{\mathrm{}}{}}}\sqrt{\left(k+1\right)\theta }|\psi _{k+1}\psi _k|.`$ (3.60) Projection into $`𝕄_{N+1}`$ gives: $$\widehat{a}_\theta ^{\left(N\right)}=\underset{s=0}{\overset{N1}{}}\sqrt{\left(s+1\right)\theta }|\psi _s\psi _{s+1}|.$$ (3.61) According to (3.59), to perform the derivative with respect to $`z`$, we have to consider: $`[\widehat{a}_\theta ^{\left(N\right)},\widehat{a}^{}]`$ $`=`$ $`\theta \left[|\psi _0\psi _0|+{\displaystyle \underset{s=1}{\overset{N1}{}}}|\psi _s\psi _s|N|\psi _N\psi _N|\right]`$ (3.62) $`=`$ $`\theta \left[{\displaystyle \underset{s=0}{\overset{N1}{}}}|\psi _s\psi _s|N|\psi _N\psi _N|\right]`$ $`=`$ $`\theta \left[{\displaystyle \underset{s=0}{\overset{N}{}}}|\psi _s\psi _s|\left(1+N\right)|\psi _N\psi _N|\right].`$ The first term of the sum is the projector onto the first $`N+1`$ basis elements, whose symbol is the identity in $`𝒜_\theta ^{\left(N\right)}`$. It is worth noting that this commutator has no terms ’outside’ the space we are considering, namely there are no components on density matrices of order greater than $`N+1`$: this means that in this case there is no need to project it on $`𝕄_{N+1}`$. The symbol yields then: $$_z\left(z_\theta ^{\left(N\right)}\right)=e^{Nr^2}\left[\underset{s=0}{\overset{N}{}}\frac{r^{2s}N^s}{s!}\frac{N+1}{N!}r^{2N}N^N\right].$$ (3.63) In the limit of $`N\mathrm{}`$ the first term is the characteristic function for the disc, while the second converges to a factor $`\pi \delta \left(r1\right)`$. This factor is a radial $`\delta `$ selecting the value $`r=1`$ with respect to the Lebesgue measure on the plane: $$\underset{N\mathrm{}}{lim}_z\left(z_\theta ^{\left(N\right)}\right)=Id\left(r\right)\pi \delta \left(r1\right).$$ (3.64) To calculate the derivative of $`\overline{z}`$ with respect to $`z`$ one needs to consider: $$\widehat{a}_\theta ^{\left(N\right)}=\underset{k=0}{\overset{N1}{}}\sqrt{\left(k+1\right)\theta }|\psi _{k+1}\psi _k|$$ and the commutator: $$[\widehat{a}_\theta ^{\left(N\right)},\widehat{a}^{}]=\theta \sqrt{N\left(N+1\right)}|\psi _{N+1}\psi _N|.$$ (3.65) Unlike the previous example this operator must be projected back to the algebra $`𝕄_{N+1}`$, and upon considering the symbol we finally have: $$_{\overline{z}}\left(z_\theta ^{\left(N\right)}\right)=0.$$ (3.66) Let us come to the definition of the Laplacian operator. In the spirit of noncommutative geometry it is this additional structure which carries the information about the geometry of the space underlying $`𝒜_\theta `$. In facts, we can say we have succeeded in fuzzifying the disc only if we have been able to define a fuzzy Laplacian whose spectrum approaches that of the ordinary Laplacian on the disc when $`N\mathrm{}`$. Starting from the exact expressions: $$^2f(\overline{z},z)=4_{\overline{z}}_zf=\frac{4}{\theta ^2}z|[\widehat{a},[\widehat{f},\widehat{a}^{}]]|z$$ (3.67) it is possible to define, in each $`𝒜_\theta ^{\left(N\right)}`$: $$_{\left(N\right)}^2\widehat{f}_\theta ^{\left(N\right)}\frac{4}{\theta ^2}\widehat{P}_\theta ^{\left(N\right)}[\widehat{a},[\widehat{P}_\theta ^{\left(N\right)}\widehat{f}\widehat{P}_\theta ^{\left(N\right)},\widehat{a}^{}]]\widehat{P}_\theta ^{\left(N\right)}.$$ (3.68) The image of the element of the truncated algebra: $$\widehat{f}_\theta ^{\left(N\right)}=\underset{a,b=0}{\overset{N}{}}f_{ab}|\psi _a\psi _b|$$ is then: $`_{\left(N\right)}^2f_\theta ^{\left(N\right)}`$ $`=`$ $`\mathrm{\hspace{0.17em}4}N[{\displaystyle \underset{s=0}{\overset{N1}{}}}{\displaystyle \underset{b=0}{\overset{N1}{}}}f_{s+1,b+1}\sqrt{\left(s+1\right)\left(b+1\right)}|\psi _s\psi _b|+`$ (3.69) $`{\displaystyle \underset{s=0}{\overset{N}{}}}{\displaystyle \underset{b=0}{\overset{N}{}}}f_{sb}\left(s+1\right)|\psi _s\psi _b|{\displaystyle \underset{s=0}{\overset{N1}{}}}f_{0,s+1}\left(s+1\right)|\psi _0\psi _{s+1}|+`$ $`+{\displaystyle \underset{s=0}{\overset{N1}{}}}{\displaystyle \underset{b=0}{\overset{N1}{}}}f_{sb}\sqrt{\left(s+1\right)\left(b+1\right)}|\psi _{s+1}\psi _{b+1}|+`$ $`{\displaystyle \underset{s=0}{\overset{N1}{}}}{\displaystyle \underset{b=0}{\overset{N1}{}}}f_{s+1,b+1}(b+1)|\psi _{s+1}\psi _{b+1}|].`$ The eigenvalues of this Laplacian can be numerically calculated. They are seen to converge to the spectrum of the continuum one for functions on a disc, with boundary conditions on the edge of the disc of Dirichlet homogeneous kind . All eigenvalues are negative, and their modules $`\lambda `$ solve the implicit equation: $$J_n\left(\sqrt{\lambda }\right)=0,$$ (3.70) where $`n`$ is the order of the Bessel functions. In particular, those related to $`J_0`$ are simply degenerate, the others are doubly degenerate: so there is a sequence of eigenvalues labelled by $`\lambda _{n,k}`$ where $`n`$ is the order of the Bessel function and $`k`$ indicates that it is the $`k^{th}`$ zero of the function. The eigenfunctions of this continuum operator are: $$\mathrm{\Phi }_{n,k}=e^{in\phi }\left(\frac{\sqrt{\lambda _{\left|n\right|,k}}r}{2}\right)^{|n|}\underset{s=0}{\overset{\mathrm{}}{}}\frac{\left(\lambda _{|n|,k}\right)^s}{s!\left(|n|+s\right)!}\left(\frac{r}{2}\right)^{2s}=e^{in\phi }J_{|n|}\left(\sqrt{\lambda _{\left|n\right|,k}}r\right).$$ (3.71) with $`n`$ integer number and $`|n|`$ its absolute value. This is a way to write the eigenfunctions in a compact form, taking into account the degeneracy of eigenvalues for $`|n|1`$. The spectrum of the fuzzy Laplacian is in good agreement with the spectrum of the continuum case, even for low values $`N`$ of the dimension of truncation, as can be seen in figure 1. It is interesting to note that with this definition of Laplacian (3.68) we correctly reproduce the pattern of non degenerate and double degenerate eigenvalues. The difference with the case of the spectrum of the fuzzy Laplacian for the fuzzy sphere is that now the ”fuzzy spectrum” is both a cut-off and an approximation of the continuum spectrum. It is a cut-off because, of course, it is a finite rank operator. It is an approximation because it has been defined using a formalism whose building blocks are related to a noncompact group, namely the Heisenberg-Weyl, whose generators have no finite dimensional realization. ## 4 Fuzzy Bessel functions In this section we introduce a basis on the fuzzy disc. As we have already noticed in the introduction, we cannot use representation theory, which in the case of the fuzzy sphere yields the fuzzy harmonics, because there is no compact group whose homogeneous space is the disc. But we have a well defined Laplacian for each $`𝒜_\theta ^{\left(N\right)}`$. Therefore we may answer the problem looking at the eigenfunctions of this fuzzy Laplacian. ### 4.1 Spectral properties of the Laplacian The fuzzy Laplacians (3.68) are a family of operators mapping each algebra $`𝒜_\theta ^{\left(N\right)}`$ into itself with the explicit action (3.69). Considering $`𝒜_\theta ^{\left(N\right)}`$ as a $`\left(N+1\right)^2`$ dimensional vector space, a density matrix $`|\psi _a\psi _b|`$ is a basis element for it, with $`a,b(0,\mathrm{},N)`$. Equation (3.69) can be specified: $`_{\left(N\right)}^2\left(|\psi _a\psi _b|\right)`$ $`=`$ $`4N\widehat{P}_\theta ^{\left(N\right)}(\sqrt{ab}|\psi _{a1}\psi _{b1}|+`$ $`(a+b+1)|\psi _a\psi _b|+\sqrt{\left(a+1\right)\left(b+1\right)}|\psi _{a+1}\psi _{b+1}|)\widehat{P}^{\left(N\right)}_\theta .`$ The action of the projector operator is to cut away some of the components of the right side image for $`a=0,N`$ and for $`b=0,N`$. It is important to note that, chosen an integer $`n=ab`$, the fuzzy Laplacian maps basis elements with the same $`n`$ onto themselves, as $`n`$ ranges through $`(N,\mathrm{},+N)`$. In the matrix representation of $`𝒜_\theta ^{\left(N\right)}`$, the number $`n`$ fixes the diagonal to which the element $`|\psi _a\psi _b|`$ belongs. The equation (4.1) then shows that the fuzzy Laplacian maps each diagonal subspace of $`𝕄_{N+1}`$ into itself. This property simplifies the study of the spectral properties of the fuzzy Laplacian. The analysis of these spectral properties will be performed restricting the operator to each of its stability subspaces. To understand the meaning of this integer $`n`$, go back to the eigenfunctions of the continuum Laplacian on a disc with Dirichlet homogeneous boundary conditions (3.71). Eigenfunctions $`\mathrm{\Phi }_{n,k}`$ are defined on the real plane, so they can be mapped into operators on the whole Hilbert space via the Weyl map introduced before (3.40). The continuum eigenfunctions (3.71) with a fixed nonnegative $`n`$, such that $`0nN`$ can be mapped into an operator: $$\widehat{\mathrm{\Phi }}_{n,k}=\left(\frac{\lambda _{n,k}}{4}\right)^{n/2}\underset{s=0}{\overset{\mathrm{}}{}}\left(\frac{\lambda _{n,k}}{4}\right)^s\frac{1}{s!\left(s+n\right)!}\widehat{a}^s\widehat{a}^{s+n}.$$ (4.2) In the density matrix notation, it acquires the form: $$\widehat{\mathrm{\Phi }}_{n,k}=\left(\frac{\lambda _{n,k}}{4}\right)^{n/2}\underset{j=n}{\overset{\mathrm{}}{}}\underset{s=0}{\overset{jn}{}}\left(\frac{\theta \lambda _{n,k}}{4}\right)^s\frac{\theta ^{n/2}}{s!\left(s+n\right)!}\frac{\sqrt{j!\left(jn\right)!}}{\left(jsn\right)!}|\psi _{jn}\psi _j|.$$ (4.3) This element can be fuzzified, so that now the parameter $`\theta `$ is constrained as $`\theta N=1`$: $$\widehat{\mathrm{\Phi }}_{n,k}^{\left(N\right)}=\left(\frac{\lambda _{n,k}}{4N}\right)^{n/2}\underset{a=0}{\overset{Nn}{}}\sqrt{a!\left(a+n\right)!}\left[\underset{s=0}{\overset{a}{}}\left(\frac{\lambda _{n,k}}{4N}\right)^s\frac{1}{s!\left(s+n\right)!\left(as\right)!}\right]|\psi _a\psi _{a+n}|.$$ (4.4) Eigenfunctions with negative $`n`$ are just the complex conjugate of those with positive $`n`$: $`\left(\mathrm{\Phi }_{n,k}\right)^{}=\mathrm{\Phi }_{n,k}`$. The fuzzification procedure gives a matrix which can be obtained by $`\widehat{\mathrm{\Phi }}_{n,k}^{\left(N\right)}`$ by an hermitian conjugation: $`\widehat{\mathrm{\Phi }}_{n,k}^{\left(N\right)}=\widehat{\mathrm{\Phi }}_{n,k}^{\left(N\right)}`$. This analysis clarifies that, in the fuzzy approximation, continuum eigenfunctions are represented by matrices belonging to the $`n^{th}`$ diagonal subspace of $`𝕄_{N+1}`$. Moreover, in this approach, admissible $`n`$ are constrained by the dimension of the fuzzification: $`nN`$. In all these fuzzy elements, $`\lambda _{n,k}`$ (as well as $`\lambda _{0,k}`$) can be seen no longer as eigenvalues of the Laplacian operator on the disc with Dirichlet homogeneous boundary conditions, but, at this stage, simply as parameters. We can now perform a complete spectral analysis of the fuzzy Laplacian. This will also show the exact meaning of the $`\lambda _{n,k}`$ parameters. Consider first the $`n=0`$ subspace, whose dimension is $`N+1`$, and consider the elements $`\widehat{\mathrm{\Phi }}_{0,k}^{\left(N\right)}`$ with $`k=1,\mathrm{}N+1`$. Act on these elements with the fuzzy Laplacian. From equation (4.1) we obtain: $$_{\left(N\right)}^2\left(|\psi _a\psi _a|\right)=4N\widehat{P}_\theta ^{\left(N\right)}\left(a|\psi _{a1}\psi _{a1}|\left(2a+1\right)|\psi _a\psi _a|+\left(a+1\right)|\psi _{a+1}\psi _{a+1}|\right)\widehat{P}_\theta ^{\left(N\right)}$$ (4.5) one can reproject the image $`_{\left(N\right)}^2\left(\widehat{\mathrm{\Phi }}_{n,k}^{\left(N\right)}\right)`$ on the basis density matrices for the specific subspace. On the first basis element, equation(4.5) becomes: $$\psi _0|_{\left(N\right)}^2\left(|\psi _a\psi _a|\right)|\psi _0=4N(a\delta _{0,a1}\left(2a+1\right)\delta _{0,a}).$$ (4.6) Considering the explicit form of (4.4) we obtain, for $`n=0`$: $$\widehat{\mathrm{\Phi }}_{0,k}^{\left(N\right)}=\underset{a=0}{\overset{N}{}}\varphi _{0,k}^{(a,N)}|\psi _a\psi _a|;$$ (4.7) where: $$\varphi _{0,k}^{(a,N)}=\left[\underset{s=0}{\overset{a}{}}\left(\frac{\lambda _{0,k}}{4N}\right)^s\frac{1}{s!}\left(\begin{array}{c}a\\ s\end{array}\right)\right],$$ (4.8) and upon applying the Laplacian: $`\psi _0|_{\left(N\right)}^2\left(\widehat{\mathrm{\Phi }}_{0,k}^{\left(N\right)}\right)|\psi _0`$ $`=`$ $`4N\left[{\displaystyle \underset{s=0}{\overset{1}{}}}{\displaystyle \frac{1}{s!}}\left(\begin{array}{c}1\\ s\end{array}\right)\left({\displaystyle \frac{\lambda _{0,k}}{4N}}\right)^s\mathrm{\hspace{0.17em}1}\right]=4N\left({\displaystyle \frac{\lambda _{0,k}}{4N}}\right)`$ (4.11) $`=`$ $`\lambda _{0,k}=\lambda _{0,k}\psi _0|\widehat{\mathrm{\Phi }}_{0,k}^{\left(N\right)}|\psi _0.`$ (4.12) Hence the fuzzy Laplacian maps the element $`\widehat{\mathrm{\Phi }}_{0,k}^{\left(N\right)}`$ into another diagonal element whose coefficient, respect to $`|\psi _0\psi _0|`$, is just a multiplication by $`\lambda _{0,k}`$ of the coefficient of $`\widehat{\mathrm{\Phi }}_{0,k}^{\left(N\right)}`$ respect to the same basis element. Proceeding along this way we project the element $`_{\left(N\right)}^2\left(\widehat{\mathrm{\Phi }}_{n,k}^{\left(N\right)}\right)`$ on $`|\psi _b\psi _b|`$, with $`b=1,\mathrm{},N1`$: $$\psi _b|_{\left(N\right)}^2\left(\widehat{\mathrm{\Phi }}_{0,k}^{\left(N\right)}\right)|\psi _b=\left(b+1\right)\left(\varphi _{0,k}^{(b+1,N)}\varphi _{0,k}^{(b,N)}\right)+b\left(\varphi _{0,k}^{(b1,N)}\varphi _{0,k}^{(b,N)}\right).$$ (4.13) To study the r.h.s. of this relation, we need to calculate the difference: $`\mathrm{\Phi }_{0,k}^{(b+1,N)}\mathrm{\Phi }_{0,k}^{(b,N)}`$ $`=`$ $`\left[{\displaystyle \underset{s=0}{\overset{b+1}{}}}\left({\displaystyle \frac{\lambda _{0,k}}{4N}}\right)^s{\displaystyle \frac{1}{s!}}\left(\begin{array}{c}b+1\\ s\end{array}\right)\right]\left[{\displaystyle \underset{s=0}{\overset{b}{}}}\left({\displaystyle \frac{\lambda _{0,k}}{4N}}\right)^s{\displaystyle \frac{1}{s!}}\left(\begin{array}{c}b\\ s\end{array}\right)\right]`$ (4.18) $`=`$ $`{\displaystyle \underset{s=1}{\overset{b+1}{}}}\left[{\displaystyle \frac{1}{s!}}\left(\begin{array}{c}b\\ s1\end{array}\right)\right]\left({\displaystyle \frac{\lambda _{0,k}}{4N}}\right)^s,`$ (4.21) hence the equation (4.13) becomes: $`\psi _b|_{\left(N\right)}^2\left(\widehat{\mathrm{\Phi }}_{0,k}^{\left(N\right)}\right)|\psi _b`$ $`=`$ $`4N\left(\left(b+1\right)\left(\varphi _{0,k}^{(b+1,N)}\varphi _{0,k}^{(b,N)}\right)+b\left(\varphi _{0,k}^{(b1,N)}\varphi _{0,k}^{(b,N)}\right)\right)`$ (4.27) $`=`$ $`4N({\displaystyle \underset{s=1}{\overset{b+1}{}}}\left[{\displaystyle \frac{b+1}{s!}}\left(\begin{array}{c}b\\ s1\end{array}\right)\right]({\displaystyle \frac{\lambda _{0,k}}{4N}})^s+`$ $`{\displaystyle \underset{s=1}{\overset{b}{}}}\left[{\displaystyle \frac{b}{s!}}\left(\begin{array}{c}b1\\ s1\end{array}\right)\right]({\displaystyle \frac{\lambda _{0,k}}{4N}})^s)`$ $`=`$ $`\lambda _{0,k}\left[{\displaystyle \underset{s=0}{\overset{b}{}}}{\displaystyle \frac{1}{s!}}\left(\begin{array}{c}b\\ s\end{array}\right)\left({\displaystyle \frac{\lambda _{0,k}}{4N}}\right)^s\right]=\lambda _{0,k}\mathrm{\Phi }_{0,k}^{(b,N)},`$ (4.30) where to reach the expression on the last line we have recollected equal powers of the $`\lambda `$ and made some simplifications. This relation is similar to (4.12). Whatever $`\lambda _{0,k}`$, the fuzzy Laplacian acts on the first $`N`$ components of the fuzzy elements $`\widehat{\mathrm{\Phi }}_{0,k}^{\left(N\right)}`$ simply as a multiplication by $`\lambda _{0,k}`$. To conclude that the elements $`\widehat{\mathrm{\Phi }}_{0,k}^{\left(N\right)}`$ are eigenstates of the fuzzy Laplacian we need to study the last component projection of $`_{\left(N\right)}^2\left(\widehat{\mathrm{\Phi }}_{0,k}^{\left(N\right)}\right)`$, namely that on the basis element $`|\psi _N\psi _N|`$. Moreover, we have not yet obtained any conditions on the possible eigenvalues of the fuzzy Laplacian. Let us project: $$\psi _N|_{\left(N\right)}^2\left(\widehat{\mathrm{\Phi }}_{0,k}^{\left(N\right)}\right)|\psi _N=4N\left(N\mathrm{\Phi }_{0,k}^{(N1,N)}\left(2N+1\right)\mathrm{\Phi }_{0,k}^{(N,N)}\right),$$ (4.31) and impose: $$\psi _N|_{\left(N\right)}^2\left(\widehat{\mathrm{\Phi }}_{0,k}^{\left(N\right)}\right)|\psi _N=\lambda _{0,k}\mathrm{\Phi }_{0,k}^{(N,N)}=\lambda _{0,k}\psi _N|\widehat{\mathrm{\Phi }}_{0,k}^{\left(N\right)}|\psi _N.$$ (4.32) This can be cast in the form: $`{\displaystyle \underset{s=0}{\overset{N}{}}}{\displaystyle \frac{1}{s!}}\left(\begin{array}{c}N\\ s\end{array}\right)\left({\displaystyle \frac{\lambda _{0,k}}{4N}}\right)^{s+1}+\left(N+1\right){\displaystyle \underset{s=0}{\overset{N}{}}}{\displaystyle \frac{1}{s!}}\left(\begin{array}{c}N\\ s\end{array}\right)\left({\displaystyle \frac{\lambda _{0,k}}{4N}}\right)^s`$ (4.37) $`+N{\displaystyle \underset{s=0}{\overset{N}{}}}{\displaystyle \frac{1}{s!}}\left(\begin{array}{c}N1\\ s1\end{array}\right)\left({\displaystyle \frac{\lambda _{0,k}}{4N}}\right)^s`$ $`=`$ $`0.`$ (4.40) Recollecting terms with equal powers of $`\lambda `$, the equation can be written as: $$E_N\left(\frac{\lambda _{0,k}}{4N}\right)\underset{a=0}{\overset{N+1}{}}\frac{\left(1\right)^{a+N+1}}{\left(N+1a\right)!}\left[\frac{\left(N+1\right)!}{a!}\right]^2\left(\frac{\lambda _{0,k}}{4N}\right)^a=0.$$ (4.41) The meaning of this calculation is clear: relation (4.32) is not identically satisfied by all value of $`\lambda _{0,k}`$. So it becomes an equation, giving the admissible eigenvalues $`\lambda _{0,k}^{\left(N\right)}`$ of the fuzzy Laplacian on the subspace of diagonal fuzzy elements in $`𝒜_\theta ^{\left(N\right)}`$: $$_{\left(N\right)}^2\left(\widehat{\mathrm{\Phi }}_{0,k}^{\left(N\right)}\right)=\lambda _{0,k}^{\left(N\right)}\widehat{\mathrm{\Phi }}_{0,k}^{\left(N\right)}$$ (4.42) if and only if: $$E_N\left(\frac{\lambda _{0,k}^{\left(N\right)}}{4N}\right)=0.$$ (4.43) We now show that the solutions of this last equation are actually the eigenvalues of the fuzzy Laplacian (in the $`n=0`$ case). From the action of $`_{(N)}^2`$ on a generic element of this diagonal subspace, for $`\widehat{\mathrm{\Phi }}^{\left(N\right)}=_{a=0}^N\varphi _a^{\left(N\right)}|\psi _a\psi _a|`$ we have: $`_{\left(N\right)}^2\widehat{\mathrm{\Phi }}^{\left(N\right)}`$ $`=`$ $`4N((\varphi _1^{\left(N\right)}\varphi _0^{\left(N\right)})|\psi _0\psi _0|+(N\varphi _{N1}^{\left(N\right)}(2N+1)\varphi _N^{\left(N\right)})|\psi _N\psi _N|+`$ (4.44) $`+{\displaystyle \underset{a=1}{\overset{N1}{}}}[(a+1)(\varphi _{a+1}^{\left(N\right)}\varphi _a^{\left(N\right)})+a(\varphi _{a1}^{\left(N\right)}\varphi _a^{\left(N\right)})]|\psi _a\psi _a|),`$ the eigenvalue equation (4.42) can be written in components as: $`\varphi _1^{\left(N\right)}\varphi _0^{\left(N\right)}`$ $`=`$ $`\left({\displaystyle \frac{\lambda _{0,k}^{\left(N\right)}}{4N}}\right)\varphi _0^{\left(N\right)}`$ $`\left(a+1\right)\left(\varphi _{a+1}^{\left(N\right)}\varphi _a^{\left(N\right)}\right)+a\left(\varphi _{a1}^{\left(N\right)}\varphi _a^{\left(N\right)}\right)`$ $`=`$ $`\left({\displaystyle \frac{\lambda _{0,k}^{\left(N\right)}}{4N}}\right)\varphi _a^{\left(N\right)}\text{if}a0,N`$ $`N\varphi _{N1}^{\left(N\right)}\left(2N+1\right)\varphi _N^{\left(N\right)}`$ $`=`$ $`\left({\displaystyle \frac{\lambda _{0,k}^{\left(N\right)}}{4N}}\right)\varphi _N^{\left(N\right)}.`$ (4.45) This set of equations can be given a matrix form. If we set $`x_N=\lambda _{0,k}^{\left(N\right)}/4N`$, then the eigenvalues are related to the roots of the characteristic polynomial $`P_N\left(x_N\right)`$, given by the determinant of the matrix $$P_N=det\left(\begin{array}{ccccc}x_N1& 1& \mathrm{}& 0& 0\\ 1& x_N3& \mathrm{}& 0& 0\\ 0& 2& \mathrm{}& 0& 0\\ \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}\\ 0& 0& \mathrm{}& x_N\left(2N1\right)& N\\ 0& 0& \mathrm{}& N& x_N\left(2N+1\right)\end{array}\right).$$ (4.46) The eigenvalue equation for the fuzzy Laplacian in the diagonal matrices subspace is: $$P_N\left(\frac{\lambda _{0,k}^{\left(N\right)}}{4N}\right)=0.$$ (4.47) Matrices of the kind of (4.46) ara called Jacobi matrix , and their characteristic polynomial is given inductively: $$P_N\left(x_N\right)=\left(x_N\left(2N+1\right)\right)P_{N1}\left(x_N\right)N^2P_{N2}\left(x_N\right).$$ (4.48) Equation (4.47) is equivalent to (4.43). This result, proven in the appendix, closes the analysis on the spectral properties for the $`n=0`$ case: eigenstates are given by $`\widehat{\mathrm{\Phi }}_{0,k}^{\left(N\right)}`$, while eigenvalues are given by the roots of (4.47). We approach the problem for the action of the fuzzy Laplacian in the nondiagonal subspaces of $`𝒜_\theta ^{\left(N\right)}`$, defined by the integer $`n`$, following a similar strategy. The dimension of a generic subspace is $`N+1n`$, the basis elements are $`|\psi _a\psi _{a+n}|`$ with $`a=0\mathrm{},Nn`$ so the fuzzy Laplacian can be seen as an operator acting on $`\text{}^{N+1n}`$. Going through the path described for the $`n=0`$ case, we prove that the eigenvalue problem in this subspace is solved by the fuzzy elements $`\widehat{\mathrm{\Phi }}_{n,k}^{\left(N\right)}`$ (4.4) with $`k=1,\mathrm{}N+1n`$: $$_{\left(N\right)}^2\widehat{\mathrm{\Phi }}_{n,k}^{\left(N\right)}=\lambda _{n,k}^{\left(N\right)}\widehat{\mathrm{\Phi }}_{n,k}^{\left(N\right)}$$ (4.49) if and only if the eigenvalues are solutions of: $$E_N^{\left(n\right)}=\underset{k=0}{\overset{Nn+1}{}}\left(1\right)^{k+Nn+1}\frac{\left(N+1\right)!\left(Nn+1\right)!}{k!\left(n+k\right)!\left(Nnk+1\right)!}\left(\frac{\lambda _{n,k}^{\left(N\right)}}{4N}\right)^k=0.$$ (4.50) This equation generalises the equation (4.41), reducing to that for $`n=0`$. The proof will closely follow the path outlined in the diagonal case. First project the image element $`_{\left(N\right)}^2\left(\widehat{\mathrm{\Phi }}_{n,k}^{\left(N\right)}\right)`$ on the first basis element $`|\psi _0\psi _n|`$. Following the general case (4.1), one can see that: $$\psi _0|\left[_{\left(N\right)}^2\left(|\psi _a\psi _{a+n}|\right)\right]|\psi _n=4N\left[\delta _{1,a}\sqrt{n+1}\left(n+1\right)\delta _{0,a}\right].$$ (4.51) After some algebra using the definition (4.4): $$\psi _0|\left[_{\left(N\right)}^2\widehat{\mathrm{\Phi }}_{n,k}^{\left(N\right)}\right]|\psi _n=\lambda _{n,k}\psi _0|\widehat{\mathrm{\Phi }}_{n,k}^{\left(N\right)}|\psi _n$$ (4.52) where the value $`\lambda _{n,k}`$ is again just a parameter. Now we can project the $`_{\left(N\right)}^2\left(\widehat{\mathrm{\Phi }}_{n,k}^{\left(N\right)}\right)`$ element on the basis elements $`|\psi _b\psi _{b+n}|`$, with $`b0,Nn`$. Again, after some straightforward algebra: $$\psi _b|\left[_{\left(N\right)}^2\widehat{\mathrm{\Phi }}_{n,k}^{\left(N\right)}\right]|\psi _{b+n}=\lambda _{n,k}\psi _b|\widehat{\mathrm{\Phi }}_{n,k}^{\left(N\right)}|\psi _{b+n}.$$ (4.53) Projecting the $`_{\left(N\right)}^2\left(\widehat{\mathrm{\Phi }}_{n,k}^{\left(N\right)}\right)`$ element on the last basis element: $`\psi _{Nn}|\left[_{\left(N\right)}^2\widehat{\mathrm{\Phi }}_{n,k}^{\left(N\right)}\right]|\psi _N=`$ (4.54) $`=4N\left({\displaystyle \frac{\lambda _{n,k}}{4N}}\right)^{n/2}({\displaystyle \underset{s=0}{\overset{Nn1}{}}}{\displaystyle \frac{\sqrt{\left(Nn\right)!N!}}{s!\left(n+s\right)!\left(Nn1s\right)!}}({\displaystyle \frac{\lambda _{n,k}}{4N}})^s+`$ $`{\displaystyle \underset{s=0}{\overset{Nn}{}}}(2Nn+1){\displaystyle \frac{\sqrt{\left(Nn\right)!N!}}{s!\left(n+s\right)!\left(Nns\right)!}}({\displaystyle \frac{\lambda _{n,k}}{4N}})^s)`$ $`.`$ This r.h.s. defined to be: $$\lambda _{n,k}\psi _{Nn}|\widehat{\mathrm{\Phi }}_{n,k}^{\left(N\right)}|\psi _N=\lambda _{n,k}\left(\frac{\lambda _{n,k}}{4N}\right)^{n/2}\left(\underset{s=0}{\overset{Nn}{}}\frac{\sqrt{\left(Nn\right)!N!}}{s!\left(n+s\right)!\left(Nns\right)!}\left(\frac{\lambda _{n,k}}{4N}\right)^s\right)$$ (4.55) This is an equation for $`\lambda _{n,k}`$, that can be cast exactly in the form of (4.50): $$E_N^{\left(n\right)}=\underset{k=0}{\overset{Nn+1}{}}\left(1\right)^{k+Nn+1}\frac{\left(N+1\right)!\left(Nn+1\right)!}{k!\left(n+k\right)!\left(Nnk+1\right)!}\left(\frac{\lambda _{n,k}}{4N}\right)^k=0.$$ (4.56) So the eigenvalues $`\lambda _{n,k}^{\left(N\right)}`$ of the fuzzy Laplacian in this stability subspace are given by the solution of: $$E_N^{\left(n\right)}\left(\frac{\lambda _{n,k}^{\left(N\right)}}{4N}\right)=0.$$ (4.57) It is again possible to prove that this equation is exactly equivalent to the secular equation coming from the matrix representation of the action of the fuzzy Laplacian on each of this $`Nn+1`$ dimensional subspaces. The analogue of the matrix (4.46) is, in this case: $$P_N^{\left(n\right)}=det\left(\begin{array}{ccccc}x_N\left(n+1\right)& \sqrt{\left(n+1\right)}& \mathrm{}& 0& 0\\ \sqrt{\left(n+1\right)}& x_N\left(n+3\right)& \mathrm{}& 0& 0\\ \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}\\ 0& 0& \mathrm{}& x_N\left(2Nn1\right)& \sqrt{N\left(Nn\right)}\\ 0& 0& \mathrm{}& \sqrt{N\left(Nn\right)}& x_N\left(2Nn+1\right)\end{array}\right).$$ (4.58) The characteristic polynomial is of degree is $`N+1n`$ and a recursive relation holds, analogous to (4.48): $$P_N^{\left(n\right)}\left(x_N\right)=\left(x_N\left(2Nn+1\right)\right)P_{N1}^{\left(n\right)}\left(x_N\right)N\left(Nn\right)P_{N2}^{\left(n\right)}\left(x_N\right).$$ (4.59) In the appendix there is the proof of the equivalence of the relation $$P_N^{\left(n\right)}\left(x_N\right)=0$$ (4.60) with (4.57). Setting $`n=0`$ the diagonal case is recovered. The analysis for the case $`n<0`$ is straightforward. The eigenvalue equation is the same, and this shows why there is a set of doubly degenerate eigenvalues. The eigenstates are obtained by those for positive $`n`$ by complex conjugation. This completes the spectral analysis of the fuzzy laplacian. Since we have proved that the eigenstates of this fuzzy Laplacian are the fuzzified version of the continuum eigenfunctions, we call $`\widehat{\mathrm{\Phi }}_{n,k}^{\left(N\right)}`$ *fuzzy Bessels*. ### 4.2 Comparison of Bessel Functions We now compare the behaviour of the symbols of the fuzzy Bessel with their ordinary counterparts. From (4.4) the symbol of a fuzzy Bessel is: $$\mathrm{\Phi }_{n,k}^{\left(N\right)}=\left(\frac{\lambda _{n,k}^{\left(N\right)}}{4}\right)^{n/2}r^ne^{in\phi }e^{Nr^2}\underset{a=0}{\overset{Nn}{}}r^{2a}N^a\left[\underset{s=0}{\overset{a}{}}\frac{1}{s!\left(s+n\right)!\left(as\right)!}\left(\frac{\lambda _{n,k}^{\left(N\right)}}{4N}\right)^s\right].$$ (4.61) The integer $`n`$ appears as a phase modulating factor for the variable $`\phi `$. This would be the expansion of the corresponding Bessel function, where it not for the truncation in the sum, and the fact that the parameter $`\lambda _{0,k}`$ has become the eigenvalue of the fuzzy Laplacian, i.e. a solution of (4.47). For the $`n=0`$ the expression can be simplified: $`\mathrm{\Phi }_{0,k}^{\left(N\right)}\left(r\right)`$ $`=`$ $`{\displaystyle \underset{a=0}{\overset{N}{}}}\left({\displaystyle \underset{s=0}{\overset{a}{}}}\left({\displaystyle \frac{\lambda _{0,k}^{\left(N\right)}}{4N}}\right)^s{\displaystyle \frac{1}{s!}}\left(\begin{array}{c}a\\ s\end{array}\right)\right)e^{Nr^2}r^{2a}{\displaystyle \frac{N^a}{a!}}`$ (4.64) $`=`$ $`e^{Nr^2}{\displaystyle \underset{a=0}{\overset{N}{}}}{\displaystyle \frac{N^a}{a!}}r^{2a}_{\left(a\right)}\left({\displaystyle \frac{\lambda _{0,k}^{\left(N\right)}}{4N}}\right).`$ (4.65) Where $`_{\left(a\right)}\left(\frac{\lambda _{0,k}^{\left(N\right)}}{4N}\right)`$ is the $`a^{th}`$ Laguerre polynomial in the variable $`\left(\frac{\lambda _{0,k}^{\left(N\right)}}{4N}\right)`$. We can plot the diagonal fuzzy elements. Fig. 2 shows that the zero order fuzzy Bessel state converges to the continuum eigenfunctions $`\mathrm{\Phi }_{0,1}(r,\phi )`$ for values of $`r`$ inside the disc of radius $`1`$, while it converges to zero outside the disc. This behaviour is seen to be valid also for eigenstates of different eigenvalues. The plots for the symbols $`\mathrm{\Phi }_{0,2}^{\left(N\right)}`$ are in figure 3, those for $`\mathrm{\Phi }_{0,3}^{\left(N\right)}`$ in figure 4. It is interesting to analyse the fuzzification of $`\mathrm{\Phi }_{0,10}\left(r\right)`$. The fuzzy symbol is $`\widehat{\mathrm{\Phi }}_{0,10}^{\left(N\right)}`$. For $`N=10,20,25`$ it is plotted in figure (5). It is evident that the fuzzy eigenfunction reproduces the continuum eigenfunction for values of $`r`$ close to the centre of the disc, but not on the edge, where a huge bump appears. In it has been explained that the presence of the bump on the edge of the disc, in the fuzzification of a function defined on the plane, is related to the fact that this function has oscillations of too small wavelenght compared to $`\theta `$. This is manifestation of the infrared-ultraviolet mixing characteristic of noncommutative theories. In the case of $`\mathrm{\Phi }_{0,10}\left(r\right)=J_0\left(\sqrt{\lambda _{0,10}}r\right)`$, one can immediately see that the oscillation wavelenght of the continuum eigenfunction is given by $`\rho _\lambda \mathrm{\hspace{0.17em}1}/k1/10`$. In these plots, it is assumed $`\theta =1/N`$ (the fuzzy disc truncation), so $`\theta `$ and $`\rho _\lambda `$ are of compatible magnitude. In the fuzzy disc limit, $`N\mathrm{}`$, so $`\theta `$ is infinitesimal. The bump disappears, as it is shown in the other plots (figure 6). The non radial functions follow a similar pattern. Their phases are exactly as the ones of their continuum counterparts, while the radial parts are similar. A first plot is in Fig. 7, a second one in Fig. 8, where the fuzzification procedure gives again a bump, for small values of $`N`$. This bump is seen to disappear in the fuzzy disc limit. The fuzzy Bessels, being eigenfunctions, provide an efficient way to calculate a fuzzy Green function, which we calculated numerically in . Since the fuzzy Laplacian is represented by an hermitian finite dimensional matrix, its inverse can be written in terms of a spectral decomposition. If a matrix, say $`M`$, is hermitian, then its components satisfy the condition $`M_{ab}=M_{ba}^{}`$ in terms of transposition and complex conjugation. The eigenvalue problem gives a number of real eigenvalues equal to the dimension of the space on which the matrix acts: $$\underset{b}{}M_{ab}v_b^{\left(k\right)}=\lambda ^{\left(k\right)}v_a^{\left(k\right)},$$ (4.66) here $`v_a^{\left(k\right)}`$ indicates the $`a^{th}`$ components of the eigenvector relative to the eigenvalue $`\lambda ^{\left(k\right)}`$. The inverse, if it exists, of the matrix $`M`$ is a matrix $`G`$ whose components can be written as: $$G_{sq}=\underset{k}{}v_s^{\left(k\right)}v_q^{\left(k\right)}/\lambda ^{\left(k\right)}.$$ (4.67) If we specialise to the problem under analysis, the notion of eigenvector of components $`v_a^{\left(N\right)}`$ with eigenvalue $`\lambda ^{\left(k\right)}`$ goes into that of fuzzy Bessel $`\widehat{\mathrm{\Phi }}_{n,k}^{\left(N\right)}`$ matrix, from which it is immediate to obtain the symbols. The fuzzy Green function becomes: $$G_\theta ^{\left(N\right)}(z,z^{})=\underset{n=N}{\overset{+N}{}}\underset{k=1}{\overset{N+1\left|n\right|}{}}\frac{\mathrm{\Phi }_{n,k}^{\left(N\right)}\left(z^{}\right)^{}\mathrm{\Phi }_{n,k}^{\left(N\right)}\left(z\right)}{\lambda _{|n|,k}^{\left(N\right)}}.$$ (4.68) Since the fuzzy Bessels play a role similar to fuzzy harmonics for the fuzzy sphere algebra, we can now make describe the process of approximating the algebra of functions on a disc with matrices more precise. In complete analogy with (2.18) and (2.19), if $`f`$ is square integrable with respect to the standard measure on the disc $`d\mathrm{\Omega }=rdrd\varphi `$, it can be expanded in terms of Bessel functions: $$f(r,\phi )=\underset{n=\mathrm{}}{\overset{+\mathrm{}}{}}\underset{k=1}{\overset{\mathrm{}}{}}f_{nk}e^{in\phi }J_{\left|n\right|}\left(\sqrt{\lambda _{\left|n\right|,k}}r\right)$$ (4.69) and it is possible to truncate: $$f^{\left(N\right)}(r,\phi )=\underset{n=N}{\overset{+N}{}}\underset{k=1}{\overset{N+1|n|}{}}f_{nk}e^{in\phi }J_{\left|n\right|}\left(\sqrt{\lambda _{\left|n\right|,k}}r\right)=\underset{n=N}{\overset{+N}{}}\underset{k=1}{\overset{N+1|n|}{}}f_{nk}\mathrm{\Phi }_{n,k}(r,\phi ).$$ (4.70) This set of functions is a vector space, but it is no more an algebra, with the standard definition of sum and pointwise product, as the product of two truncated functions will get out of the algebra. The mapping from truncated functions into finite rank matrices: $$f^{(N)}\widehat{f}_\theta ^{\left(N\right)}=\underset{n=N}{\overset{+N}{}}\underset{k=1}{\overset{N+1|n|}{}}f_{nk}\widehat{\mathrm{\Phi }}_{n,k}^{\left(N\right)}$$ (4.71) endows the set of functions with a noncommutative product, inherited from the matrix product, which makes it into a non abelian algebra. The formal limit $`N\mathrm{}`$ with the constraint $`N\theta =1`$ is the abelian algebra of functions on the disc. The sequence of nonabelian algebras $`𝒜_\theta ^{(N)}`$ is what we call the *fuzzy disc*. ## A Weyl-Wigner Formalism In this appendix we give a very short introduction to the Weyl-Wigner formalism used to define maps between functions on a “classical” plane and a set of operators on a Hilbert space . Given a symplectic real, finite dimensional, vector space $`(L,\omega )`$, where the symplectic form $`\omega `$ is translationally invariant, i.e. with constant coefficients, the set of Displacement operators define a Weyl system, a unitary representation of canonical commutation relations in the exponentiated unitary version : $`\widehat{D}`$ $`:`$ $`L𝒰\left(\right)`$ $`\widehat{D}\left(z+u\right)`$ $`=`$ $`e^{i\omega (z,u)/2\theta }\widehat{D}\left(z\right)\widehat{D}\left(u\right).`$ (A.1) Here $`\theta `$ is a parameter: the Displacement operators can be seen to set a unitary projective representation of the translation group, where the phase factor is related to the sympletic structure defined on the vector space $`L`$. These operators also define a representation of the Heisenberg-Weyl group . On every one dimensional subspace of $`L`$ ($`\alpha `$ is a real scalar), via the Stone-von Neumann theorem: $$\widehat{D}\left(\alpha z\right)=e^{i\alpha \widehat{G}\left(z\right)\theta }$$ (A.2) and these generators satisfy: $$[\widehat{G}\left(z\right),\widehat{G}\left(u\right)]=i\theta \omega (z,u)\mathrm{\hspace{0.17em}1}1.$$ (A.3) This shows that a Weyl system is a way to formalise a set of canonical commutation relations for quantum observables, recovered as hermitian generators of a unitary ray representation. In this perspective $`\theta `$ acquires the role of a noncommutativity parameter. In the standard approach, the symplectic vector space is seen as a phase space for the classical dynamics of point particle: generators of a Weyl system represent the position and momentum observable for a quantum dynamics of point particle. Using the definition of a Weyl system, it is possible to define a Weyl map, that is a map from functions on the vector space $`\left(L\omega \right)`$ (dim $`L=2n`$) to $`Op\left(\right)`$, the set of operators on the Hilbert space on which Displacements are realised: $$\widehat{f}=\frac{dw}{\left(2\pi \theta \right)^n}\stackrel{~}{f}\left(w\right)\widehat{D}\left(w\right),$$ (A.4) where: $$\stackrel{~}{f}\left(w\right)=\frac{dz}{\left(2\pi \theta \right)^n}f\left(z\right)e^{i\omega (z,w)/\theta }$$ (A.5) is the symplectic Fourier transform of the function $`f`$. The Weyl map can be seen as a quantization map for classical observables defined on a phase space: it can be inverted by the Wigner map: $$\stackrel{~}{f}\left(w\right)=Tr\left[\widehat{f}\widehat{D}^{}\left(w\right)\right].$$ (A.6) So the Weyl-Wigner maps define a bijection between a set of classical observables and a set of quantum observables. This is the building block of a formalism suited to study both the problem of quantizing a classical system and performing a classical limit of a quantum system. The standard Weyl-Wigner maps can be modified by the introduction of a term, called weight: $`\widehat{f}_\lambda `$ $`=`$ $`{\displaystyle \frac{dw}{\left(2\pi \theta \right)^n}\stackrel{~}{f}\left(w\right)\lambda \left(w\right)\widehat{D}\left(w\right)}`$ $`\stackrel{~}{f}\left(w\right)`$ $`=`$ $`\lambda \left(w\right)^1Tr\left[\widehat{f}_\lambda \widehat{D}^{}\left(w\right)\right].`$ (A.7) This weight can be proved to be related to a more general class of ray representation of the translation group, and is seen to be responsible both for a specific ”ordering” in the quantization procedure and a definition of a specific domain of mathematical applicability of the bijection, as stressed in the main text. ## B Generalised coherent states In this paper the concept of generalised coherent states has been extensively used. The aim of this appendix is to briefly recollect the main definitions and results, to fix notations and facilitate the reading. The main reference for coherent states is . Consider a Lie group $`G`$, and $`\widehat{U}\left(g\right)`$ a unitary irreducible representation of this group on a Hilbert space $``$. Chosen a *fiducial* vector $`|\psi _0`$ in $``$, one obtains a set of vectors for each element of the group, acting on it with $`\widehat{U}\left(g\right)`$: $$|\psi _g\widehat{U}\left(g\right)|\psi _0.$$ (B.1) Two such vectors are considered *equivalent* if they correspond, quantum-mechanically, to the same state, i.e. if they differ by a phase. So $`|\psi _g|\psi _g^{}`$ if $`|\psi _g=e^{i\varphi (g,g^{})}|\psi _g^{}`$. This condition is equivalent to $`\widehat{U}\left(g^1g\right)|\psi _0=e^{i\varphi (g,g^{})}|\psi _0`$. If $`H`$ is the subgroup of $`G`$ whose elements are represented, by $`\widehat{U}`$, as operators whose action on the fiducial vector is just a multiplication by a phase, then the equivalence relation is among points of $`G`$, and the quotient is the space $`G/H`$. If $`H`$ is maximal, then it is called isotropy subgroup for the state $`|\psi _0`$. Choosing a representative $`g\left(x\right)`$ in each equivalence class $`xX=G/H`$ (which is a cross section of the fiber bundle $`G`$ with base $`X`$) defines a set of vectors on $``$, depending, clearly, on $`G`$ and $`|\psi _0`$. This set of states is called a *system of coherent states* for $`G`$. The state corresponding to the vector $`|x`$ may be considered as the range of a rank one projector in $``$. Thus, the system of generalised coherent states determines a set of one dimensional subspaces in $``$, parametrised by points of the homogeneous space $`X=G/H`$. An evolution of this analysis drives naturally to the issue of overcompleteness for the system of coherent states, mentioned in (3.34), (2.6). ## C Explicit Calculations for the Eigenvalues The aim of this appendix is to show an explicit proof of the results claimed in the main text about the eigenvalue problem for the fuzzy Laplacian. In particular, we want to prove that the equation (4.57) is equivalent to the equation (4.60). From this equivalence, it will be proved also the equivalence between equations (4.43) and equation (4.47), as they are a special case of the first two for $`n=0`$. The polynomial $`E_N^{\left(n\right)}`$ is given, identifying $`\lambda _{n,k}^{\left(N\right)}/4N=x^N`$, by (4.50): $$E_N^{\left(n\right)}\left(x_N\right)=\underset{k=0}{\overset{Nn+1}{}}\left(1\right)^{k+Nn+1}\frac{\left(N+1\right)!\left(Nn+1\right)!}{k!\left(n+k\right)!\left(Nnk+1\right)!}\left(x_N\right)^k=0.$$ (C.2) The characteristic polynomial $`P_N^{\left(n\right)}`$ is given by the determinant of the matrix (4.58) given recursively by (4.59). To prove the equality of the two polynomials, the first step is to prove that $`E_N^{\left(n\right)}`$ satisfies the same recursive relation that $`P_N^{\left(n\right)}`$ does. Considering the quantity: $`\left(x_N\left(2Nn+1\right)\right)E_{N1}^{\left(n\right)}\left(x_N\right)N\left(Nn\right)E_{N2}^{\left(n\right)}\left(x_N\right)=`$ $`={\displaystyle \underset{k=0}{\overset{Nn}{}}}\left(1\right)^{k+Nn}{\displaystyle \frac{N!\left(Nn\right)!}{k!\left(n+k\right)!\left(Nnk\right)!}}\left(x_N\right)^{k+1}+`$ $`{\displaystyle \underset{k=0}{\overset{Nn}{}}}\left(1\right)^{k+Nn}\left(2Nn+1\right){\displaystyle \frac{N!\left(Nn\right)!}{k!\left(n+k\right)!\left(Nnk\right)!}}\left(x_N\right)^k+`$ $`{\displaystyle \underset{k=0}{\overset{Nn1}{}}}\left(1\right)^{k+Nn+1}{\displaystyle \frac{N!\left(Nn\right)!}{k!\left(n+k\right)!\left(Nnk1\right)!}}\left(x_N\right)^k.`$ (C.3) Where the coefficient of the zeroth order term is: $$\left(\left(1\right)^{Nn+1}\frac{N!}{n!}\left(2Nn+1\right)\left(1\right)^{Nn+1}\frac{N!}{n!}\left(Nn\right)\right)=\left(1\right)^{Nn+1}\frac{\left(N+1\right)!}{n!}.$$ (C.4) It coincides with the coefficient of the zeroth order term of the polynomial $`E_N^{\left(n\right)}`$. Let us consider, in the expression (C.3), the coefficient of the $`q^{th}`$ order term, with $`1qNn1`$. It is equal to: $`(1)^{Nn+q1}N!(Nn)!({\displaystyle \frac{1}{\left(a1\right)!\left(n+a1\right)!\left(Nna+1\right)!}}+`$ $`{\displaystyle \frac{2Nn+1}{a!\left(n+a\right)!\left(Nna\right)!}}{\displaystyle \frac{1}{a!\left(n+a\right)!\left(Nna1\right)!}}\left)\right(x_N)^q=(1\left)^{Nn+a+1}{\displaystyle \frac{\left(N+1\right)!\left(Nn+1\right)!}{a!\left(a+n\right)!\left(Nna+1\right)!}}\right(x_N)^q`$ $`.`$ This coefficient is equal to the coefficient of the $`q^{th}`$ order term in $`E_N^{\left(n\right)}`$. The equality of the coefficients, of the remaining order terms, of the polynomial (C.3) with those of the polynomial $`E_N^{\left(n\right)}\left(x_N\right)`$ can be easily checked. At this stage, we have proved that the polynomial $`E_N^{\left(n\right)}\left(x_N\right)`$ satisfies the same recursive relation that the polynomial $`P_N^{\left(n\right)}\left(x_N\right)`$ does satisfy. The recursive relation they satisfy is such that the $`N^{th}`$ term of the sequence $`E_N^{\left(n\right)}`$, for each fixed $`n`$, depends on both the $`\left(N1\right)^{th}`$ and $`\left(N2\right)^{th}`$ terms of the sequence. To prove the complete equivalence of these two polynomials, we thus need to prove that they explicitly coincide at a pair of consecutive steps. In the main text it has been stressed that there is a constraint for allowed positive $`n`$, namely an upper bound: $`nN+1`$. This means that, for a fixed $`n`$, that is for a fixed subspace in the fuzzy algebra, $`Nn+1`$. For $`N=n1`$ both polynomials $`E_{n1}^{\left(n\right)}\left(x_N\right)`$ and $`P_{n1}^{\left(n\right)}\left(x_N\right)`$ are trivial, so we study at first the case of $`N=n`$. We have: $$E_n^{\left(n\right)}=\underset{k=0}{\overset{1}{}}\left(1\right)^{k+1}\frac{\left(n+1\right)!}{k!\left(n+k\right)!\left(1k\right)!}\left(x^{\left(N\right)}\right)^k=x^{\left(N\right)}\left(n+1\right)$$ (C.6) while the matrix representing the action of the fuzzy Laplacian is one-dimensional, so that: $$P_n^{\left(N\right)}=x_N\left(2n+1\right).$$ (C.7) The second case is for $`N=n+1`$. We have: $`E_n^{\left(n+1\right)}`$ $`=`$ $`{\displaystyle \underset{k=0}{\overset{2}{}}}\left(1\right)^{k+2}{\displaystyle \frac{2\left(n+2\right)!}{k!\left(n+k\right)!\left(2k\right)!}}\left(x_N\right)^k`$ (C.8) $`=`$ $`\left(x_N\right)^22\left(n+2\right)x_N+\left(n+1\right)\left(n+2\right).`$ The matrix representing the action of the fuzzy Laplacian is now two-dimensional, so that the characteristic polynomial is: $`P_{n+1}^{\left(n\right)}`$ $`=`$ $`det\left(\begin{array}{cc}x_N\left(n+1\right)& \sqrt{n+1}\\ \sqrt{n+1}& x_N\left(n+3\right)\end{array}\right)=`$ (C.11) $`=`$ $`\left(x_N\right)^22\left(n+2\right)x_N+\left(n+1\right)\left(n+2\right).`$ (C.12) This proves the equivalence of the two polynomials. ### Acknowledgments We thank D. Bercioux, P. Lucignano, P. Santorelli and A. Stern for help at various stages. This work has been supported in part by the Progetto di Ricerca di Interesse Nazionale SInteSi.
warning/0506/astro-ph0506240.html
ar5iv
text
# High energy 𝛾-ray emission from the starburst nucleus of NGC 253 ## 1 Introduction Starburst galaxies (and in general star forming regions) are expected to be detected as $`\gamma `$-ray sources. They have both a large amount of target material and, due to the presence of supernova remnants and the powerful stellar winds of their numerous young stars, multiple shocks where to accelerate particles up to relativistic energies. Pion decay production of $`\gamma `$-rays, usually dominant under such conditions, is thought to produce significant $`\gamma `$-ray fluxes. Approximately 90% of the high-energy $`\gamma `$-ray luminosity of the Milky Way ($``$1.3 $`\times 10^6`$ L, Strong, Moskalenko, & Reimer 2000) is diffuse emission from cosmic ray interactions with interstellar gas and photons (Hunter et al. 1997). However, the LMC is the only external galaxy that has been detected in its diffuse $`\gamma `$-ray emission (Sreekumar et al. 1992), a fact explained by the isotropic flux dilution by distance. At 1 Mpc, for example, the flux of the Milky Way would approximately be 2.5 $`\times 10^8`$ photons cm<sup>-2</sup> s<sup>-1</sup> ($`>`$100 MeV), what is below the sensitivity achieved up to now in the relevant energy domain. The Energetic $`\gamma `$-ray Experiment, EGRET, did not detect any starburst. Upper limits were imposed for M82, $`F(E>100\mathrm{M}\mathrm{e}\mathrm{V})<4.4\times 10^8`$ photons cm<sup>-2</sup> s<sup>-1</sup>, and NGC 253, $`F(E>100\mathrm{M}\mathrm{e}\mathrm{V})<3.4\times 10^8`$ photons cm<sup>-2</sup> s<sup>-1</sup> (Blom et al. 1999), the two nearest starbursts, as well as upon many luminous infrared galaxies, for which similar constraints were found (Torres et al. 2004). These upper limits are barely above the theoretical predictions of models for the $`\gamma `$-ray emission of galaxies, constructed with different levels of detail (see Torres 2004b for a review). Starbursts and luminous infrared galaxies are expected to compensate the flux dilution produced by their relatively larger distance to Earth with their higher cosmic ray flux, and become sources for the next generation of instruments measuring photons in the 100 MeV – 10 GeV regime, like the Gamma-ray Large Area Telescope, GLAST (e.g., Völk, Aharonian & Breitschwerdt 1996, Paglione et al. 1996, Blom et al. 1999, Torres et al. 2004, Torres 2004). In this paper, we analyze one such galaxy, the well studied starburst NGC 253. Herein, we present a self-consistent multiwavelength modelling of the central region of the galaxy, taking into account the latest measurements of densities, supernova explosion rate, radio emission, and molecular mass, among other physical parameters that we use as input for our scenario. ## 2 CANGAROO observations of NGC 253 Recently, the CANGAROO collaboration reported the detection of NGC 253 at TeV $`\gamma `$-ray energies (an 11$`\sigma `$ confidence level claim), observed during a period of two years in 2000 and 2001 by about 150 hours (Itoh et al. 2002, 2003). Their measured differential flux was fitted by Itoh et al. (2003) with a power law and an exponential cutoff, obtaining $`{\displaystyle \frac{dF}{dE}}=(2.85\pm 0.71)\times 10^{12}(E/1\mathrm{T}\mathrm{e}\mathrm{V})^{(3.85\pm 0.46)}\mathrm{cm}^2\mathrm{s}^1\mathrm{TeV}^1,`$ and $`{\displaystyle \frac{dF}{dE}}=ae^{\sqrt{E_0}/b}(E/E_0)^{1.5}e^{\sqrt{E}/b}\mathrm{cm}^2\mathrm{s}^1\mathrm{TeV}^1,`$ with $`a=6\times 10^5\mathrm{cm}^2\mathrm{s}^1\mathrm{TeV}^1,E_0=0.0002\mathrm{TeV},`$ and $`b=0.25\pm 0.01\sqrt{\mathrm{TeV}}.`$ Both parameterizations are sensible reproductions of the observational data, although the former is clearly preferred for simplicity upon the light of an equally good fit. The flux uncertainty are the square root of the quadratic sum of the statistical and systematic errors. Note that the slope of the power law spectrum is very uncertain, but steep. Indeed, an extrapolation of this power-law spectrum to lower energies violates the measured upper limits in the GeV regime. The CANGAROO collaboration suggested a turn-over below the TeV region and proposed the second spectral form. They have also claimed that the emission at the highest energies is inconsistent with it being produced in a point like source, and proposed for it an inverse Compton origin in a kpc-scale $`\gamma `$-ray halo. The HESS array have observed NGC 253 (see below). In several other observations of sources that have previously been targets for CANGAROO, the HESS collaboration have presented results in clear contradiction with the former CANGAROO reports. This is most notably the case for SN 1006 (Aharonian et al. 2005a), PSR B1706-44 (Aharonian et al. 2005b), and to some extent also for the supernova remnant RX J1713-3857 (Aharonian et al. 2004a), and the Galactic Center (Aharonian et al. 2004b). This may suggest some kind of systematic difference in the treatment of both sets of observational data. Such systematic effect should explain why CANGAROO spectra are steeper and their measured fluxes are one order of magnitude higher than the upper limits or measurements obtained with HESS. The CANGAROO collaboration is now calibrating their stereo system, and will be re-observing these problematic cases within a year or so (R. Enomoto 2005, private communication). In what follows we focus on producing a detailed multiwavelength theoretical model for the central region of NGC 253, irrespective of CANGAROO measurements (i.e., we will not try to reproduce their spectrum, but we will derive predictions of fluxes based on a set of well-founded assumptions). The aim is to see whether a model based on observations at all wavelengths and first principles would –while being consistent with multiwavelength testing– predict that the central region of NGC 253 alone can produce a sufficiently high $`\gamma `$-ray flux so as to be detected by the current ground-based Cerenkov telescopes and future MeV–GeV satellites. The central region of the galaxy would look like a point like source for the field of view and angular resolution of imaging Cerenkov telescopes. Then, we shall explore if we would rather theoretically expect a HESS non-confirmation of CANGAROO results regarding perhaps both, the flux and the extension. ## 3 Phenomenology of the central region of NGC 253 A wealth of new multiwavelength data was obtained for NGC 253 during the last decade (after the previous modelling by Paglione et al. 1996, which did not include photons energies above 100 GeV, see below). It is located at a distance of $`2.5`$ Mpc (Turner and Ho 1985, Maurbersger et al. 1996) and it is a nearly edge-on (inclination 78<sup>o</sup>, Pence 1981) barred Sc galaxy. The continuum spectrum of NGC 253 peaks in the FIR at about 100 $`\mu `$m, with a luminosity of $`4\times 10^{10}`$ L (Telesco & Harper 1980, Rice et al. 1988, Melo et al. 2002). The FIR luminosity is at least a factor of 2 larger than that of our own Galaxy (Cox & Mezger 1989, Dudley & Wynn-Williams 1999), and it mainly (about half of it according to Melo et al.’s (2002) 1 arcmin resolution ISOPHOT observations) comes from the central nucleus. IR emission can be understood as cold ($`T50K`$) dust reprocessing of stellar photon fields. When observed at 1 pc resolution, at least 64 individual compact radio sources have been detected within the central 200 pc of the galaxy (Ulvestad & Antonucci 1997), and roughly 15 of them are within the central arcsec of the strongest radio source, considered to be either a buried active nucleus or a very compact SNR. Of the strongest 17 sources, about half have flat spectra and half have steep spectra. This indicates that perhaps half of the individual radio sources are dominated by thermal emission from H II regions, and half are optically thin synchrotron sources, presumably SNRs. There is no compelling evidence for any sort of variability in any of the compact sources over an 8 yr time baseline. The most powerful flat-spectrum central radio source is clearly resolved in the study of Ulvestad and Antonucci (1997) and appears to be larger than the R136 cluster located in 30 Doradus, containing about 10<sup>5</sup> M in stars and 600 M in ionized gas. The age was estimated to be less than 4 $`\times `$ 10<sup>6</sup> yr. The region surrounding the central 200 pc has also been observed with subarcsec resolution and 22 additional radio sources stronger than 0.4 mJy were detected within 2kpc of the galaxy nucleus (Ulvestad 2000). The region outside the central starburst may account for about 20% of the star formation of NGC 253, is subject to a supernova explosion rate well below $`0.1`$ yr<sup>-1</sup>, and has an average gas density in the range 20–200 cm<sup>-3</sup>, much less than in the most active nuclear region of NGC 253 (Ulvestad 2000). Carilli (1996) presented low frequency radio continuum observations of the nucleus at high spatial resolution. Free-free absorption was claimed to be the mechanism producing a flattening of the synchrotron curve at low energies, with a turnover frequency located between 10<sup>8.5</sup> and 10<sup>9</sup> Hz. The emission measures needed for this turnover to happen, for temperatures in the order of 10<sup>4</sup> K, is at least 10<sup>5</sup> pc cm<sup>-6</sup>. Tingay (2004) observed NGC 253 using the Australian Long Baseline Array and provided fits with free-free absorption models for the radio spectrum of six sources. He concluded that the free-free opacity in the central region has to be in the range of 1 to 4 at 1.4 GHz, implying emission measures of a few times 10<sup>6</sup> pc cm<sup>-6</sup> in this particular directions, again for temperatures of the order of 10<sup>4</sup> K. As shown by infrared, millimeter, and centimeter observations, the 200 pc central region dominates the current star formation in NGC 253, and is considered as the starburst central nucleus (e.g., Ulvestad and Antonucci 1997, Ulvestad 2000). Centimeter imaging of this inner starburst, and the limits on variability of radio sources, indicates a supernova rate less than 0.3 yr<sup>-1</sup> (Ulvestad & Antonucci 1997), which is consistent with results ranging from 0.1 to 0.3 yr<sup>-1</sup> inferred from models of the infrared emission of the entire galaxy (Rieke et al. 1980; Rieke, Lebofsky & Walker 1988, Forbes et al. 1993). Van Buren and Greenhouse (1994) developed, starting from Chevalier’s (1982) model for radio emission from supernovae blast waves expanding into the ejecta of their precursor stars, a direct relationship between the FIR luminosity and the rate of supernova explosions. The result is $`=2.3\times 10^{12}L_{\mathrm{FIR}}/L_{}`$ yr<sup>-1</sup>, which is in agreement, within uncertainties, with the previous estimates. The star formation rate at the central region has been computed from IR observations, resulting in 3.5 M yr<sup>-1</sup>, and represents about 70% of the total star formation rate measured for NGC 253 (Melo et al. 2002). When compared with Local Group Galaxies, the supernova rate in NGC 253 is one order of magnitude larger than that of M31, the largest of the Local Group (Pavlidou and Fields 2001). Paglione et al. (2004) obtained high resolution (5”.2 $`\times `$ 5”.2) interferometric observations of the CO line $`J=10`$ in order to study the structure and kinematics of the molecular gas in the central nucleus. This study enhances that of Sorai et al. (2000), which, although done with less angular resolution, obtained compatible results. The general morphology of the CO map is consistent with other high resolution studies. It shows an extended ridge of emission aligned with an infrared-bright bar and a central group of bright clouds aligned with the major axis of the galaxy, orbiting the radio nucleus in a possible ring. The central clouds move around the radio nucleus as a solid body, similar to the distribution of the radio sources, central HCN clouds, and central near-infrared emission (Paglione et al. 1995, 1997; Ulvestad & Antonucci 1997). Much of the molecular gas in NGC 253 appears to be highly excited (Wild et al. 1992; Mao et al. 2000; Ward et al. 2003). Observations of $`J=43`$ and $`J=65`$ transitions of CO as well as HCN lines suggest the existence of localized spots with values of densities in excess of 10<sup>4</sup> cm<sup>-3</sup> (Israel & Baas 2002, Paglione, Jackson, & Ishizuki 1997, Paglione, Tosaki & Jackson 1995, Harris et al. 1991). Bradford et al. (2003) have reported CO $`J=76`$ observations and also find that the bulk of molecular gas in the central 180 pc of the galaxy is highly excited and at a temperature of about 120 K. They concluded that the best mechanism for heating the gas is cosmic ray bombardment over the gas residing in clouds, with density about 4.5 $`\times 10^4`$ cm<sup>-3</sup>. Current estimates of the gas mass in the central 20” – 50” ($`<600`$ pc) region range from 2.5 $`\times 10^7`$ M (Harrison, Henkel & Russell 1999) to 4.8 $`\times 10^8`$ M (Houghton et al. 1997), see Bradford et al. (2003), Sorai et al. 2000, and Engelbracht et al. (1998) for discussions. For example, using the standard CO to gas mass conversion, the central molecular mass was estimated as 1.8 $`\times 10^8`$ M (Mauersberger et al. 1996). It would be factor of $`3`$ lower if such is the correction to the conversion factor in starburst regions which are better described as a filled intercloud medium, as in the case of ULIRGs, instead of a collection of separate large molecular clouds, see Solomon et al. (1997), Downes & Solomon (1998), and Bryant & Scoville (1999) for discussions. Thus we will assume in agreement with the mentioned measurements that within the central 200 pc, a disk of 70 pc height has $``$ 2 $`\times 10^7`$ M uniformly distributed, with a density of $`600`$ cm<sup>-3</sup>. Additional target gas mass with an average density of $``$50 cm<sup>-3</sup> is assumed to populate the central kpc outside the innermost region, but subject to a smaller supernova explosion rate $`0.01`$ yr<sup>-1</sup>, 10% of that found in the most powerful nucleus (Ulvestad 2000). The central region of this starburst is packed with massive stars. Watson et al. (1996) have discovered four young globular clusters near the center of NGC 253; they alone can account for a mass well in excess of 1.5$`\times 10^6`$M (see also Keto et al. 1999). Assuming that the star formation rate has been continuous in the central region for the last 10<sup>9</sup> yrs, and a Salpeter IMF for 0.08–100 M, Watson et al. (1996) find that the bolometric luminosity of NGC 253 is consistent with 1.5 $`\times 10^8`$M of young stars. Physical, morphological, and kinematic evidence for the existence of a galactic superwind has been found for NGC 253 (e.g., McCarthy et al. 1987, Heckman et al. 1990, Strickland et al. 2000, 2002, Pietsch et al. 2001, Forbes et al. 2000, Weaver et al. 2002, Sugai, Davies & Ward 2003). This superwind creates a cavity of hot ($`10^8`$ K) gas, with cooling times longer than the typical expansion timescales. As the cavity expands, a strong shock front is formed on the contact surface with the cool interstellar medium. Shock interactions with low and high density clouds can produce X-ray continuum and optical line emission, respectively, both of which have been directly observed (McCarthy et al. 1987). The shock velocity can reach thousands of km s<sup>-1</sup>. This wind has been proposed as the convector of particles which have been already accelerated in individual SNRs, to the larger superwind region, where Fermi processes could upgrade their energy up to that detected in ultra high energy cosmic rays (Anchordoqui et al. 1999, Anchordoqui et al. 2003, Torres & Anchordoqui 2004). ## 4 Diffuse modelling The approach to compute the steady multiwavelength emission from NGC 253 follows that implemented in $`𝒬`$-diffuse, which we have used with some further improvements (Torres 2004). The steady state particle distribution is computed within $`𝒬`$-diffuse as the result of an injection distribution being subject to losses and secondary production in the ISM. In general, the injection distribution may be defined to a lesser degree of uncertainty when compared with the steady state one, since the former can be directly linked to observations. The injection proton emissivity, following Bell (1978), is assumed to be a power law in proton kinetic energies, with index $`p`$, $`Q_{\mathrm{inj}}(E_{\mathrm{p},\mathrm{kin}})=K(E_{\mathrm{p},\mathrm{kin}}/\mathrm{GeV})^p`$ , where $`K`$ is a normalization constant and units are such that $`[Q]`$= GeV<sup>-1</sup> cm<sup>-3</sup> s<sup>-1</sup>. The normalization $`K`$ is obtained from the total power transferred by supernovae into CRs kinetic energy within a given volume $`_{E_{\mathrm{p},\mathrm{kin},\mathrm{min}}}^{E_{\mathrm{p},\mathrm{kin},\mathrm{max}}}Q_{\mathrm{inj}}(E_{\mathrm{p},\mathrm{kin}})E_{\mathrm{p},\mathrm{kin}}𝑑E_{\mathrm{p},\mathrm{kin}}=KE_{\mathrm{p},\mathrm{kin},\mathrm{min}}^{p+2}/(p+2)_i\eta _i𝒫_i/V`$ where it was assumed that $`p2`$, and used the fact that $`E_{\mathrm{p},\mathrm{kin}\mathrm{min}}E_{\mathrm{p},\mathrm{kin},\mathrm{max}}`$ in the second equality. $`_i`$ ($`_i_i=`$) is defined as the rate of supernova explosions in the star forming region being considered, $`V`$ being its volume, and $`\eta _i`$, the transferred fraction of the supernova explosion power ($`𝒫10^{51}`$ erg) into CRs (e.g., Torres et al. 2003 and references therein). The summation over $`i`$ takes into account that not all supernovae will transfer the same amount of power into CRs (alternatively, that not all supernovae will release the same power). The rate of power transfer is assumed to be in the range 0.05 $`\eta _i`$ 0.15, the average value for $`\eta `$ being $`10\%`$. At low energies the distribution of cosmic rays is flatter, e.g., it would be given by equation (6) of Bell (1978), correspondingly normalized. We have numerically verified that to neglect this difference at low energy does not produce any important change in the computation of secondaries, and especially on $`\gamma `$-rays at the energies of interest. The general diffusion-loss equation is given by (see, e.g., Longair 1994, p. 279; Ginzburg & Syrovatskii 1964, p. 296) $$D^2N(E)+\frac{N(E)}{\tau (E)}\frac{d}{dE}\left[b(E)N(E)\right]Q(E)=\frac{N(E)}{t}$$ (1) In this equation, $`D`$ is the scalar diffusion coefficient, $`Q(E)`$ represents the source term appropriate to the production of particles with energy $`E`$, $`\tau (E)`$ stands for the confinement timescale, $`N(E)`$ is the distribution of particles with energies in the range $`E`$ and $`E+dE`$ per unit volume, and $`b(E)=\left(dE/dt\right)`$ is the rate of loss of energy. The functions $`b(E)`$, $`\tau (E)`$, and $`Q(E)`$ depend on the kind of particles. In the steady state, $`N(E)/t=0,`$ and, under the assumption of a homogeneous distribution of sources, the spatial dependence is considered to be irrelevant, so that $`D^2N(E)=0.`$ Eq. (1) can be solved by using the Green function $$G(E,E^{})=\frac{1}{b(E)}\mathrm{exp}\left(_E^E^{}𝑑y\frac{1}{\tau (y)b(y)}\right),$$ (2) such that for any given source function, or emissivity, $`Q(E)`$, the solution is $$N(E)=_E^{E_{\mathrm{max}}}𝑑E^{}Q(E^{})G(E,E^{}).$$ (3) The confinement timescale will be determined by several contributions. One on hand, we consider the characteristic escape time in the homogeneous diffusion model (Berezinskii et al. 1990, p. 50-52 and 78) $`\tau _D=R^2/(2D(E))=\tau _0/(\beta (E/\mathrm{GeV})^\mu ),`$ where $`\beta `$ is the velocity of the particle in units of $`c`$, $`R`$ is the spatial extent of the region from where particles diffuse away, and $`D(E)`$ is the energy-dependent diffusion coefficient, whose dependence is assumed $`E^\mu `$, with $`\mu 0.5`$.<sup>1</sup><sup>1</sup>1 We emphasize that the use of an homogeneous model is an assumption, but justified in the compactness of the innermost starburst region. We are basically assuming that there is an homogeneous distribution of supernovae in the central hundreds of pc, what is supported observationally (Ulvestad and Antonucci 1997). $`\tau _0`$ is the characteristic diffusive escape time at $``$ 1 GeV. $`\tau _0`$ for NGC253 is not known. One can only assume its value and compare it with that for other galaxies (e.g. our own Galaxy, or M33, Duric et al. 1995); the value we choose also parallels that obtained in an earlier study of NGC 253 or on M82 (Paglione et al. 1996, Blom et al. 1999). We analyze the sensitivity of the model to $`\tau _0`$ below. On the other, the total escape timescale will also take into account that particles can be carried away by the collective effect of stellar winds and supernovae. $`\tau _c`$, the convective timescale, is $`R/V`$, where $`V`$ is the collective wind velocity. For a wind velocity of 300 km/s and a radius of about the size of the innermost starburst (see below), this timescale is less than a million years ( 3$`\times 10^5`$ yr). The outflow velocity is not very well known, however, but minimum reasonable values between 300 and 600 km s<sup>-1</sup> have been claimed, and could even reach values of the order of thousand of km s<sup>-1</sup> (Strickland et al. 2002). Pion losses (which are catastrophic, since the inelasticity of the collision is about 50%) produce a loss timescale $`\tau _{\mathrm{pp}}^1=(dE/dt)^{\mathrm{pion}}/E`$ (see, e.g., Mannheim & Schlickeiser 1994), which is similar in magnitude to the convective timescale and dominates with it the shaping of the proton spectrum. Thus, in general, for energies higher than the pion production threshold $`\tau ^1(E)=\tau _D^1+\tau _{c}^{}{}_{}{}^{1}+\tau _{\mathrm{pp}}^1.`$ For electrons, the total rate of energy loss considered is given by the sum of that involving ionization, inverse Compton scattering, bremsstrahlung, and synchrotron radiation. We have also incorporated adiabatic losses. Full Klein-Nishina cross section is used while computing photon emission, and either Thomson or extreme Klein-Nishina approximations, as needed, are used while computing losses. For the production of secondary electrons, $`𝒬`$-diffuse computes knock-on electrons and charged pion processes producing both electrons and positrons. In the case of $`\gamma `$-ray photons, we compute bremsstrahlung, inverse Compton and neutral pion decay processes. For the latter, an Appendix provides a more detailed discussion of the different approaches to compute the neutral pion-induced $`\gamma `$-ray emissivity. For radio photons, we compute, using the steady distribution of electrons, the synchrotron, and free-free emission. Free-free absorption is also considered in order to reproduce the radio data at low frequencies. The FIR emission is modelled with a dust emissivity law given by $`\nu ^\sigma B(ϵ,T)`$, where $`\sigma =1.5`$ and $`B`$ is the Planck function. The computed FIR photon density is used as a target for inverse Compton process as well as to give account of losses in the $`\gamma `$-ray scape. The latter basically comes from the opacity to $`\gamma \gamma `$ pair production with the photon field of the galaxy nucleus. The fact that the dust within the starburst reprocesses the UV star radiation to the less energetic infrared photons implies that the opacity to $`\gamma \gamma `$ process is significant only at the highest energies. The opacity to pair production from the interaction of a $`\gamma `$-ray photon in the presence of a nucleus of charge $`Z`$ is considered too. For further details and relevant formulae see Torres (2004). ### 4.1 Comparison with previous models When compared with the previous study on high energy emission from NGC 253, by Paglione et al. (1996), several methodological and modelling differences are to be mentioned. Paglione et al.’s assumed distance, size, gas mass, density, and supernova explosion rate of the central region are different from those quoted in Table 1. The former authors modelled, based on earlier data (e.g., Canzian et al. 1988) a starburst region at 3.4 Mpc (a factor of 1.36 farther than the one currently adopted), of 325 pc radius (about 3 times larger than the one adopted here). This region is larger than what is implied by the current knowledge of the central starburst, where the supernova explosion rate Paglione et al. used is actually found and the cosmic ray density is maximally enhanced. The average density assumed by Paglione et al., 300 cm<sup>-3</sup>, gives a target mass $`2\times 10^8`$ M, which is at the upper end of all current claims for the central nucleus, or already found as excessive. The target mass of the innermost region differs from ours by a factor of about 6, ours being smaller. The fraction of the supernova explosion converted into cosmic rays (20% for Paglione et al., a factor of 2 larger than ours) seems also excessive in regards of the current measurements of SNR at the highest energies. We have also considered, especially to test its influence in the total $`\gamma `$-ray output, a surrounding disk with a smaller supernova rate, following the discovery of several SNRs in that region (Ulvestad 2000). Finally, Paglione et al. (1996) study did not produce results above 200 GeV. <sup>2</sup><sup>2</sup>2To further ease the comparison, we here note some typos in Paglione et al. (1996) paper: The factor $`b(E)`$ should be elevated to the minus one in their equation (4), as well as the term $`\tau _c`$ in their equation (3). The y-axis of their Figure 1 is not the emissivity, but the emissivity divided by the density; units need to be accordingly modified, see e.g. Abraham et al. (1966). $`E_p`$ in their equation (7) and $`x`$-axis of Figure 2 and 3 is the kinetic energy, but the generic $`E`$ in Equation (1) is the total energy. The $`y`$-axis of Figure 2 is in units of cm<sup>-3</sup> GeV<sup>-1</sup>. The $`𝒬`$-diffuse set uses different parameterizations for pion cross sections as compared with those used by Marscher and Brown (1978), whose code was the basis of Paglione et al.’s study. Our computation of neutral pion decay $`\gamma `$-rays is discussed in the Appendix. In addition, $`𝒬`$-diffuse uses the full inverse Compton Klein-Nishina cross section, computes secondaries (e.g., knock-on electrons) without resorting to parameterizations which are valid only for Earth-like cosmic ray (CR) intensities, fixes the photon target for Compton scattering starting from modelling of the observations in the FIR, and considers opacities to $`\gamma `$-ray scape. ## 5 Results ### 5.1 Summary of model parameters We begin the discussion of our results by presenting a summary of the parameters we have used for, and obtained from, our modelling. These are given in Table 1. There, the mark OM refers to Obtained from modelling and ST or see text refers to parameters discussed in more detail in the previous Section on phenomenology, where references are also given. These parameters values or ranges of values are fixed by observations. Finally, the mark A refers to assumed parameters, in general within a range. Variations to the values given in Table 1 are discussed below. ### 5.2 Steady proton and electron population The numerical solution of the diffusion-loss equation for protons and electrons, each subject to the losses previously described, is shown in Figure 1. We have adopted a diffusive residence timescale of 10 Myr, a convective timescale of 1 Myr, and a density of $`600`$ cm<sup>-3</sup>. In the case of electrons, the magnetic field with which synchrotron losses are computed in Figure 1 is 300 $`\mu `$G. The latter is fixed below, requiring that the steady electron population produces a flux level of radio emission matching observations. An injection electron spectrum is considered –in addition to the secondaries– in generating the steady electron distribution. The primary electron spectrum is assumed as that of the protons times a scaling factor; the inverse of the ratio between the number of protons and electrons, $`N_p/N_e`$ (e.g., Bell 1978). This ratio is about 100 for the Galaxy, but could be smaller in star forming regions, where there are multiple acceleration sites. For instance, Völk et al. (1989) obtained $`N_p/N_e30`$ for M82. $`N_p/N_e=50`$ is assumed for the central disk of NGC 253. From about $`E_em_e10^1`$ to $`10`$ GeV, the secondary population of electrons (slightly) dominates, in any case. This is shown in Figure 3. It is in this region of energies where most of the synchrotron radio emission is generated, and thus the ability of producing a high energy model compatible with radio observations is a cross check for the primary proton distribution. Figure 2 shows the ratio of the steady proton population in the SD to that in the IS. Because the SD is subject to a smaller supernova explosion rate, it has an smaller number of protons in its steady distribution, at all energies, of the order of 1% of that of the IS. Then, it will play a subdominant role in the generation of $`\gamma `$-ray emission, as we show below. In the right panel of Figure 2, and for further discussion in the following Sections, we present the ratio between the steady proton distribution in the IS, when the gas density is artificially enhanced and diminished by a factor of 2 from the assumed value of 600 cm<sup>-3</sup>. ### 5.3 IR and radio emission The continuum emission from NGC 253, at wavelengths between $`1`$ cm and $`10`$ microns, was measured by several authors, e.g., see Figure 4 and Section 3. These observations did not in general distinguish, due to angular resolution, only the emission coming from the innermost starburst region. Instead, they also contain a contribution coming from photons produced in the surrounding disk and farther away from the nucleus. The IR continuum emission is mainly produced thermally, by dust, and thus it could be modelled with a spectrum having a dilute blackbody (graybody) emissivity law, proportional to $`\nu ^\sigma B(ϵ,T)`$, where $`B`$ is the Planck function. Figure 4 shows the result of this modelling and its agreement with observational data when the dust emissivity index $`\sigma =1.5`$ and the dust temperature $`T_{\mathrm{dust}}=50K`$. Different total luminosities, the normalization of the dust emission (see the appendix of Torres 2004 for details), are shown in the Figure to give an idea of the contribution of the innermost region with respect to that of the rest of the galaxy. According to Melo et al. (2002), about half of the total IR luminosity is produced in the IS, what is consistent with the data points being intermediate between the curves with $`L_{\mathrm{IR}}`$ 2$`\times 10^{10}`$ and 4$`\times 10^{10}`$ L, since the latter were obtained with beamsizes of about 20–50 arcsec ($``$ 240–600 pc at the NGC 253 distance). The influence of the magnetic field upon the steady state electron distribution is twofold. On one hand, the greater the field, the larger the synchrotron losses –what is particularly visible at high energies, where synchrotron losses play a relevant role. On the other, the larger the field the smaller the steady distribution. These effects evidently compete between each other in determining the final radio flux. The magnetic field is required to be such that the radio emission generated by the steady electron distribution is in agreement with the observational radio data. This is achieved by iterating the feedback between the choice of magnetic field, the determination of the steady distribution, and the computation of radio flux, additionally taking into account free-free emission and absorption processes. Whereas free-free emission is subdominant when compared with the synchrotron flux density, free-free absorption plays a key role at low frequencies, determining the opacity. We have found a reasonable agreement with all observational data for a magnetic field in the innermost region of 300 $`\mu `$G, an ionized gas temperature of about 10<sup>4</sup> K, and an emission measure of $`5\times 10^5`$ pc cm<sup>-6</sup>, the latter two are in separate agreement with the free-free modelling of the opacity of particular radio sources, as discussed in Section 3. The value of magnetic field we have found for the IS is very similar to that found for the disk of Arp 220 (Torres 2004) and compatible with measurements in molecular clouds (Crutcher 1988, 1994, 1999). ### 5.4 $`\gamma `$-ray emission In the left panel of Figure 5, bremsstrahlung, inverse Compton, and pion decay $`\gamma `$-ray fluxes of the IS are shown together with the total contribution of the SD and the total differential flux of the whole system. These results are obtained with the model just shown to be in agreement with radio and IR observations. As mentioned before, the contribution of the SD, even when having a factor of $`8`$ more mass than the IS, is subdominant. The reason for this needs to be found in that this region is much less active (Ulvestad 2000). Our predictions, while complying with EGRET upper limits, are barely below them. If this model is correct, NGC 253 is bound to be a bright $`\gamma `$-ray source for GLAST. The integral fluxes are shown in the right panel of Figure 5. Our model complies again with the integral EGRET upper limit for photons above 100 MeV, and predicts that, given enough observation time, NGC 253 is also to appear as a point-like source in an instrument like HESS. Note, however, that quite a long exposure may be needed to detect the galaxy, and also, that our fluxes are only a few percent of those reported by the CANGAROO collaboration. An additional source of TeV photons not considered here is the hadronic production in the winds of massive stars (Romero & Torres 2003). However, a strong star forming region such as the nucleus of NGC 253 is bound to possess much more free gas than that contained within the winds of massive stars, which albeit numerous, have mass loss rates typically in the range of 10<sup>-6</sup>–10<sup>-7</sup> M.<sup>3</sup><sup>3</sup>3In Romero & Torres (2003), higher mass loss rates up to $`10^5`$ M, i.e., grammages between 50 and 150 g cm<sup>-2</sup> were allowed. These values, although have been found in perhaps one or two Galactic early O stars, are uncommon. Since the size of the base of the wind for each star, the grammage, and the ambient enhancement of cosmic rays were independently allowed to take values within their assumed ranges in the Monte Carlo simulation of Romero & Torres (2003), the stars with the most favorable parameters would dominate the sum, overestimating the relative importance of their fluxes. ### 5.5 Opacities to $`\gamma `$-ray escape Two sources of opacities are considered: pair production from the $`\gamma `$-rays interaction with the photon field or with the charged nucleus present in the medium. Compton scattering and attenuation in the magnetic field by one-photon pair production are negligible. The opacity to $`\gamma \gamma `$ pair production with the photon field, which, at the same time, is target for inverse Compton processes, can be computed as $`\tau (R_c,E_\gamma )^{\gamma \gamma }=_{R_c}^{\mathrm{}}n(ϵ)\sigma _{e^{}e^+}(ϵ,E_\gamma )^{\gamma \gamma }𝑑r𝑑ϵ<\tau (E_\gamma )_{\mathrm{max}}^{\gamma \gamma }<(h/\mathrm{cos}(i))_0^{\mathrm{}}n(ϵ)\sigma _{e^{}e^+}(ϵ,E_\gamma )𝑑ϵ`$ where $`ϵ`$ is the energy of the target photons, $`E_\gamma `$ is the energy of the $`\gamma `$-ray in consideration, $`R_c`$ is the place where the $`\gamma `$-ray photon was created within the system, and $`\sigma _{e^{}e^+}(ϵ,E_\gamma )^{\gamma \gamma }=(3\sigma _T/16)(1\beta ^2)(2\beta (\beta ^22)+(3\beta ^4)\mathrm{ln}((1+\beta )/(1\beta ))),`$ with $`\beta =(1(mc^2)^2/(ϵE_\gamma ))^{1/2}`$ and $`\sigma _T`$ being the Thomson cross section, is the cross section for $`\gamma \gamma `$ pair production (e.g. Cox 1999, p.214). Note that the lower limit of the integral on $`ϵ`$ in the expression for the opacity is determined from the condition that the center of mass energy of the two colliding photons should be such that $`\beta >0`$. The fact that the dust within the starburst reprocesses the UV star radiation to the less energetic infrared photons implies that the opacities to $`\gamma \gamma `$ process is significant only at the highest energies. No source of this kind of opacity is assumed outside the system under consideration, since the nearness of NGC 253 makes negligible the opacity generated by cosmological fields. The cross section for pair production from the interaction of a $`\gamma `$-ray photon in the presence of a nucleus of charge $`Z`$ in the completely screened regime ($`E_\gamma /mc^21/(\alpha Z)`$) is independent of energy, and is given by (e.g. Cox 1999, p.213) $`\sigma _{e^{}e^+}^{\gamma Z}=(3\alpha Z^2\sigma _T/2\pi )(7/9\mathrm{ln}(183/Z^{1/3})1/54)`$. At lower energies the relevant cross section is that of the no-screening case, which has a logarithmic dependency on energy, $`\sigma _{e^{}e^+}^{\gamma Z}=(3\alpha Z^2\sigma _T/2\pi )(7/9\mathrm{ln}(2E_\gamma /mc^2)109/54)`$, and matches the complete screening cross section at around 0.5 GeV. Both of these expression are used to compute the opacity, depending on $`E_\gamma `$. In the left panel of Figure 6 we show the different contributions to the opacity. The equation of radiation transport appropriate for a disk is used to compute the predicted $`\gamma `$-ray flux taking into account all absorption processes (see Appendix of Torres 2004 for details). The right panel of Figure 6 shows the effect of the opacity on the integral $`\gamma `$-ray fluxes, only evident above 3 TeV. ### 5.6 Exploring the parameter space and degeneracies As it is summarized in Table 1, most of the model parameters are well fixed from observations. There are however a couple of assumptions which, whereas being not well bounded from experiments, may produce slight degeneracies. Consider for instance the proton injection slope $`p`$ and the diffusive scale $`\tau _0`$. For the former we have assumed $`p=2.3`$, which is in agreement with the recent results from HESS regarding $`\gamma `$-ray observations at TeV energies of supernova remnants and unidentified extended sources. However, it would be certainly within what one would expect from proton acceleration in supernova remnants, and also within experimental uncertainty, if a better description for the average proton injection slope in NGC 253 is 2.2 instead of 2.3. Table 2 shows the influence of this kind of choice on our final results. A harder slope slightly increases the integral flux. Similarly, the diffusive timescale is not well determined, and it may be arguable perhaps within one order of magnitude. Table 2 also shows the influence of this choice. Ultimately, high energy $`\gamma `$-ray observations (from GeV to TeV) are the ones to impose constraints on these values. In any case, we remark that pp interaction and convection timescales are much shorter ($`<1`$ Myr) and thus dominate the form of $`N(E)`$. To show this in greater detail we show in Figure 7 the result for the proton distribution when different convective and the pp timescales are taken into account as compared with the solution when $`\tau (E)=\tau _D`$, i.e., diffusion only. Clearly, convection plus pp timescales dominates the spectrum. Regarding degeneracies, both the proton slope and the confinement timescales, however, cannot be much different from what we have assumed. If the former were to differ significantly, it would be impossible to reproduce the radio data, which is the result of the synchrotron emission of the secondary electrons. Changes in the number of protons in the IS would imply a change in the magnetic field to reproduce radio observations, what clearly cannot be pushed much either. In a less impacting way, varying the value of $`N_p/N_e`$ can also vary the results. This variation would be slight because of the influence of the more numerous secondary electrons in the energetic region of interest for radio emission. On the same track, varying the diffusion coefficient $`\mu `$ does not import substantial changes. And if the maximum proton energy were to differ from the value of 100 TeV we have assumed (what we do not expect it to happen in a significant way, since we do now observationally know that supernova remnants are sources of $`10`$ TeV photons), the end of the spectrum would slightly shift accordingly. Even within an artificially enlarged uncertainty of the gas density, the results will not be modified much: if for any reason the average particle density were to be a factor of 2 smaller or larger, the $`\gamma `$-ray integral flux variations would be within 4% for energies above 100 MeV, and within 25% for energies above 200 GeV. Table 3 shows these results by presenting the integral fluxes above a given threshold if the assumed density of 600 cm<sup>-3</sup> is doubled or halved. As can be seen in the right panel of Figure 2, if the density is larger (smaller) by a factor $`2`$, the resultant steady proton distribution from the same proton injection population is smaller (bigger) by a similar factor over a wide range of proton energies. As $`\gamma `$-ray emissivities are proportional to both the medium density and the number of steady protons, the variations in $`\gamma `$-ray fluxes are greatly compensated. ### 5.7 Energetics and cosmic ray enhancement The left panel of Figure 8 presents the energy density contained in the steady proton population above a certain energy, i.e., based on Figure 1, the curve shows the integral $`_EN_p(E_p)E_p𝑑E_p`$. The total energy density contained by the steady population of cosmic rays above 1 GeV is about $`10^3`$ of the power emitted by all supernova explosions in the last 5 million years. The energy density contained in the steady electron population is orders of magnitude less important. The cosmic ray enhancement is a useful parameter in estimations of $`\gamma `$-ray luminosities in different scenarios. It is defined as the increase in the cosmic ray energy density with respect to the local value, $`\omega _{\mathrm{CR},}(E)=_EN_p(E_p)E_p𝑑E_p`$, where $`N_p`$ is the local cosmic ray distribution obtained from the measured cosmic ray flux that we quote in the Appendix. The enhancement factor $`\varsigma `$ is then a function of energy given by $`\varsigma (E)=(_EN_p(E_p)E_p𝑑E_p)/\omega _{\mathrm{CR},}(E)`$. Values of enhancement for NGC 253 were proposed $`\varsigma 3000`$ for energies above 1 GeV (e.g., Suchkov et al. 1993), and we can actually verify this in our model. The right panel of Figure 8 presents the enhancement factor as a function of proton energy. The larger the energy, the larger the enhancement, due to the steep decline ($`E^{2.75}`$) of the local cosmic ray spectrum. ## 6 HESS observations The HESS array has just released (Aharonian et al. 2005c) their results for NGC 253. These are based on data taken during the construction of the array with 2 and 3 telescopes operating. The total observation time was 28 hs, with a mean zenith angle of about 14 degrees. Only events where at least two telescopes were triggered were used, to enable stereoscopic reconstruction. The energy threshold for this dataset was 190 GeV. Upper limits from H.E.S.S. on the integral flux of $`\gamma `$-rays from NGC 253 (99 % confidence level) are shown, together with our predictions, in the right panel of Figure 5, and zoomed in the region above 100 GeV in Figure 9. As an example, above 300 GeV, the upper limit is $`1.9\times 10^{12}`$ photons cm<sup>-2</sup> s<sup>-1</sup>. It can be seen that our predictions are below these upper limits at all energies but still above HESS sensitivity for reasonable observation times. ## 7 Concluding remarks We have presented a multifrequency model of the central region of NGC 253. Following recent observations, we have modelled an innermost starburst with a radius of 100 pc and a supernova explosion rate of 0.08 yr<sup>-1</sup>, and a surrounding disk up to a 1 kpc in radius with an explosion rate about tenfold smaller. As a result of our modelling we have found that a magnetic field of 300 $`\mu `$G for the innermost region is consistent with high resolution radio observations, with the radiation at 1 GHz being mostly produced by secondary electrons of cosmic ray interactions. The magnetic field found for the innermost part of NGC 253 is typical of dense molecular clouds in our Galaxy, and is close to the (270 $`\mu `$G) value proposed by Weaver et al. (2002) using the equipartition argument. We have estimated free-free emission and absorption, and considered opacities to the $`\gamma `$-ray escape. The hard X-ray emission from IC and bremsstrahlung processes produced in this model is below observational constraints, e.g., by OSSE, in agreement with previous estimations of bremsstrahlung diffuse emission (Bhattacharya et al. 1994). This is consistent with measurements in the center of the Galaxy, where INTEGRAL have shown that hard X-ray emission is not diffusively, but produced by point like sources (Lebrun et al. 2004). The flux predicted is based on a set of a few well founded assumptions, mainly a) that supernova remnants accelerate most of the cosmic rays in the central region of NGC 253, and b) that they interact with the present gas, whose amount has been measured using a variety of techniques. The low opacity to $`\gamma `$-ray escape secure that basically all $`\gamma `$-rays produced in the direction towards Earth reach us. Observational constraints establishes the values of the supernova explosion rate and gas content (see Section 2 for references). The ease of all the assumptions made in our model, its concurrence with all observational constraints, and the unavoidability of the processes analyzed, lead us to conclude that 1) GLAST will detect NGC 253, being our predicted luminosity ($`2.3\times 10^8`$ photons cm<sup>-2</sup> s<sup>-1</sup> above 100 MeV) well above its 1 yr all sky survey sensitivity (GLAST Science Requirements Doc. 2003); 2) that our predicted TeV fluxes are about one order of magnitude smaller than what was claimed by CANGAROO, and thus, that perhaps in this case, a similar problem to that found in other sources affected their data taking or analysis, and 3) that HESS could detect the galaxy as a point like source provided it is observed long enough with the full array ($`50`$ hours, for a detection between 300 and 1000 GeV.) <sup>4</sup><sup>4</sup>4HESS site latitude provides that NGC 253 can be observed very close to the zenith (the minimum zenith angle for NGC 253 from HESS site is 2 degrees). As a consequence, HESS observations of NGC 253 can be done with the minimum energy threshold of the experiment. The MAGIC Telescope, although being at a northern hemisphere site, is also able to observe NGC 253 at a larger zenith angle, about 53 degrees. We finally note that this model predicts a steady $`\gamma `$-ray source, so that a posteriori variability estimators (e.g., Torres et al. 2001) can be checked for consistency. ## Appendix: Parameterizations of proton-proton cross sections for neutral pion decay The $`\pi ^0`$ emissivity resulting from an isotropic intensity of protons, $`J_p(E_p)`$, interacting with fixed target nuclei with number density $`n`$, through the reaction $`p+pp+p+\pi ^0p+p+2\gamma `$, is given by (e.g., Stecker 1971) $$Q_{\pi ^0}(E_{\pi ^0})=4\pi n_{E_{th}(E_{\pi ^0})}^{E_p^{max}}𝑑E_pJ_p(E_p)\frac{d\sigma (E_{\pi ^0},E_p)}{dE_{\pi ^0}},$$ (4) where $`E_p^{max}`$ is the maximum energy of protons in the system, and $`E_{th}(E_{\pi ^0})`$ is the minimum proton energy required to produce a pion with total energy $`E_{\pi ^0}`$, and is determined through kinematical considerations. It is obtained using the invariant, $`\sqrt{s}=\left(2m_p(E_p+m_p)\right)^{1/2}=(M_X^2+E_{\pi }^{}{}_{}{}^{2}m_\pi ^2)^{1/2}+E_{\pi }^{}{}_{}{}^{2}`$, where $`s`$ is the square of the total energy in the center-of-mass system, $`M_X`$ depends on the reaction channel and represents the invariant mass of the system consisting of all particles except the pion, $`E_\pi ^{}`$ is the CMS energy of the produced meson (that is connected with the laboratory system energy via a Lorentz transformation, see Appendices of Moskalenko & Strong 1998 and Blattnig et al. 2000b). $`E_\pi ^{}=(sM_x^2+m_\pi ^2)/(2\sqrt{s})`$, so that for a given value of $`s`$, $`E_\pi ^{}`$ will be maximum when $`M_x`$ takes its minimum value. For the case of neutral pion production $`M_x=2m_p`$. The laboratory system pion energy, obtained from $`E_\pi ^{}`$, can be put as a function of $`s`$. Inverting this relation, thus obtaining $`s=s(E_\pi )`$, and use of $`s=s(E_\pi )=2m_p(E_p+m_p)`$ allow for the minimum proton energy to be derived. Finally, $`d\sigma (E_{\pi ^0},E_p)/dE_{\pi ^0}`$ is the differential cross section for the production of a pion with energy $`E_{\pi ^0}`$, in the lab frame, due to a collision of a proton of energy $`E_p`$ with a hydrogen atom at rest. The $`\gamma `$-ray emissivity is obtained from the neutral pion emissivity $`Q_{\pi ^0}`$ as $$Q_\gamma (E_\gamma )_\pi =2_{E_{\pi ^0}^{min}(E_\gamma )}^{E_{\pi ^0}^{max}(E_p^{max})}𝑑E_{\pi ^0}\frac{Q_{\pi ^0}(E_{\pi ^0})}{(E_{\pi ^0}^2m_{\pi ^0}^2c^4)^{1/2}}$$ (5) where $`E_{\pi ^0}^{min}(E_\gamma )=E_\gamma +m_{\pi ^0}^2c^4/(4E_\gamma )`$ is the minimum pion energy required to produce a photon of energy $`E_\gamma `$ (e.g., Stecker 1971), and $`E_{\pi ^0}^{max}(E_p^{max})`$ is the maximum pion energy that the population of protons can produce. It is obtained using the invariant equation for the maximum proton energy resident in the system. Thus, an accurate knowledge of the differential cross section for pion production becomes very important to estimate the $`\gamma `$-ray emissivity. Note that $`d\sigma (E_{\pi ^0},E_p)/dE_{\pi ^0}`$ can be thought as containing the inclusive total inelastic cross section (i.e. the cross section multiplied by the average pion multiplicity). This can be stated explicitly as done, for instance, in Dermer’s (1986a) equation 3. ### $`\delta `$-function approximation In this formalism (Aharonian and Atoyan 2000), $`Q_{\pi ^0}(E_{\pi ^0})=4\pi n{\displaystyle _{E_{th}(E_{\pi ^0})}}𝑑E_pJ_p(E_p)\delta (E_{\pi ^0}\kappa E_{\mathrm{kin}})\sigma (E_p),`$ $`={\displaystyle \frac{4\pi n}{\kappa }}J_p\left(m_pc^2+{\displaystyle \frac{E_{\pi ^0}}{\kappa }}\right)\sigma \left(m_pc^2+{\displaystyle \frac{E_{\pi ^0}}{\kappa }}\right)`$ (6) where $`\sigma `$ is the total cross-section of inelastic pp collisions, and $`\kappa `$ is the mean fraction of the kinetic energy $`E_{\mathrm{kin}}=E_pm_pc^2`$ of the proton transferred to the secondary meson per collision. In a broad region from GeV to TeV energies, $`\kappa 0.17`$. In this approximation, then, an accurate knowledge of the total inelastic cross section is needed to compute the $`\gamma `$-ray emissivity. Aharonian and Atoyan (2000) proposed that, since from the threshold at $`E_{\mathrm{kin}}`$ 0.3 GeV the cross section appears to rise rapidly to about 30 mb at energies about $`E_{\mathrm{kin}}`$ 2 GeV, and since after that energy it increases only logarithmically, a sufficiently good approximation is to assume $`\sigma `$ $``$ $`30(0.95+0.06\mathrm{ln}(E_{\mathrm{kin}}/\mathrm{GeV}))\mathrm{mb}\mathrm{for}E>1\mathrm{G}\mathrm{e}\mathrm{V}`$ (7) $``$ $`0\mathrm{otherwise}.`$ It can be seen (e.g., Dermer 1986a) that different parameterizations of the cross section below 1 GeV do not noticeably affect the results of $`\gamma `$-ray emissitivities, since most of the $`\gamma `$-rays are generated by primary protons having more energy than a few GeV, provided the spectrum of primaries is sufficiently broad. In a recent paper, Kamae et al. (2005) introduced the effect of diffractive interactions and scaling violations in $`pp\pi ^0`$ interactions. The diffractive interactions contribution was usually neglected in all other computations of $`\gamma `$-ray emissivity from neutral pion decay to date, and thus one would expect an increase in the predicted fluxes. Kamae et al.’s best model, dubbed A, for the inelastic (not inclusive) cross section is given in Table 1 of their paper, columns 2 and 3. When one compares the sum of both diffractive and non-diffractive contributions of Kamae et al.’s model with the Aharonian and Atoyan’s formula, one sees that the latter produces an actually larger (but quite close) cross section. Figure 10 shows these results above proton kinetic energies of 1 GeV, as well as the difference between these cross sections. Kamae et al.’s model A was compared with Hagiwara’s (2002) compilation of pp cross section measurements and found in good agreement. When multiplicity is taken into account, Kamae et al.’s model also agrees with the data on inclusive cross sections, a point we discuss in more detail below. Figure 11 shows a comparison of the $`\gamma `$-ray emissivity obtained when using Kamae et al.’s model A and equation (6). Curves are practically indistinguishable in this scale, and their ratio is well within a factor of $`1.3`$. In this comparison, the proton spectrum is the Earth-like one, $`J_p(E_p)=2.2E_p^{2.75}`$ protons cm<sup>-2</sup> s<sup>-1</sup> sr<sup>-1</sup> GeV<sup>-1</sup> and $`n=1`$ cm<sup>-3</sup>. The resulting $`\gamma `$-ray emissivity is multiplied by 1.45 to give account of the contribution to the pion spectrum produced in interactions with heavier nuclei both as targets and as projectiles (Dermer 1986a). Aharonian and Atoyan’s expression for the cross section produces a slightly larger value of emissivity than that obtained with Kamae’s model A, including non-diffractive interactions. ### Differential cross sections parameterizations Recently, Blattnig et al. (2000) developed parameterizations of the differential cross sections. They have presented a parameterization of the Stephens and Badhwar’s (1981) model by numerically integrating the Lorentz-invariant differential cross section (LIDCS). The expression of such parameterization is divided into two regions, depending on the (laboratory frame) proton energy (Blattnig et al. 2000, see their equations 23 and 24). Blattnig et al. have also developed an alternative parameterization that has a much simpler analytical form. It is given by $`{\displaystyle \frac{d\sigma (E_{\pi ^0},E_p)}{dE_{\pi ^0}}}=e^A\mathrm{mb}\mathrm{GeV}^1`$ (8) $`\mathrm{with}`$ $`A=\left(5.8{\displaystyle \frac{1.82}{(E_pm_p)^{0.4}}}+{\displaystyle \frac{13.5}{(E_{\pi ^0}m_{\pi ^0})^{0.2}}}{\displaystyle \frac{4.5}{(E_{\pi ^0}m_{\pi ^0})^{0.4}}}\right)`$ Both parameterizations were integrated and compared with experimental results up to $`50`$ GeV in Blattnig et al.’s (2000) paper. It was found that a single expression is needed to represent the total inclusive cross section, $$\sigma _{\pi ^0}(E_p)=(0.007+0.1\frac{\mathrm{ln}(E_pm_p)}{(E_pm_p)}+\frac{0.3}{(E_pm_p)^2})^1\mathrm{mb},$$ (9) where rest masses and energies must be given in units of GeV. These two differential cross section parameterizations proposed by Blattnig et al. are not deprived of problems if extrapolated to high energy. The parameterization of the Stephen and Badhwar’s model grossly underpredicts, whereas the newest Blattnig et al. (equation 8) overpredicts, the highest energy pion yield. The $`\gamma `$-ray photon yield that is output of the use of these two differential cross sections in Equations (4,5) is also shown, for an Earth like spectrum, in Figure 11. However, the inclusive total inelastic cross section (9) seems to work well at energies higher than 50 GeV, what we show in Figure 12 together with a compilation of experimental data (Dermer 1986b). We also show in the same Figure the results for the inclusive total cross section from model A of Kamae et al. (2005), obtained from his figure 5. Indeed, Kamae et al.’s model produces a slightly higher cross section, although both correlate reasonably well with experimental data, at least up to 3 TeV.<sup>5</sup><sup>5</sup>5This figure enlarge the comparison of Blattnig et al. (2000) inclusive cross section with experimental data (see their figure 4), where only three low-energy data points from Whitmore (1974) were considered. It is the differential cross section parameterization (given by Equation 8) the one that looks suspicious at such high energies. Figure 12, right panel, shows the emissivities as computed in the different approaches multiplied by $`E^{2.75}`$. As expected, Stephen and Badhwar’s parameterization falls quickly at high energy whereas both Kamae et al.’s and Aharonian and Atoyan’s cross sections secure that the $`\gamma `$-rays emitted maintain an spectrum close to that of the proton primaries. For $`\gamma `$-rays above few TeV, i.e., $`\gamma `$-rays mostly generated by protons above few tens of TeV, Blattnig et al. differential cross section parameterization makes the $`\gamma `$-ray emitted spectrum much harder than the proton spectrum that produced them. This signals that a direct extrapolation of Eq. (8) for computing photon emissivity above TeV with Eq. (5) induces overpredictions of fluxes. Table 4 presents the results for the integrated emissivity, $`_EQ(E)𝑑E`$, with $`Q(E)`$ being the different curves of Figure 11. To obtain integrated fluxes from a source of volume $`V`$ at a distance $`D`$ one has to multiply by the constant $`V/(4\pi D^2)`$, so that the difference in integrated emissivities indeed represent those among integrated fluxes. As Table 4 shows –disregarding those coming from the Blattnig et al.’s parameterization of Stephen and Badwhar’s results which are quoted here just for completeness– above 100 MeV, differences are less than a factor of 1.5, which most likely is washed away by other uncertainties in any given model. But above 300 GeV, difference are larger and a conservative choice is in order. If interested in the GLAST-domain (say, $`E>100`$ MeV, $`E<50`$ GeV) predictions, the most conservative choice seems to be the use of the Blattnig et al. new differential cross section parameterization (Eq. 8), with no other approximation, in Equations (4) and (5). This choice, while not taking into account non-diffractive processes will possibly slightly underpredict the integrated flux (as shown in Table 4). Up to this moment, there is no public parameterization of the differential cross section including diffractive effects, but one is to be presented soon (T. Kamae, private communication). By using Blattnig et al. approach, there is no $`\delta `$-function approximation involved nor an ad-hoc histogram of particle numbers as proposed in the treatment of Kamae et al. (2005). One has an analytical expression that can be directly used in the numerical estimates of Eq. (5). However, the price to pay is that this form of computation cannot be considered reliable at higher energies and should not be used. For the IACTs-domain ($`E>100`$ GeV) the safest and also computationally-preferable choice appears to be to take either Kamae et al.’s model A, or even the simpler Aharonian and Atoyan’s expression (Eq. 7) and a $`\delta `$-function approximation. This approach would probably be slightly underestimating the integrated flux at such high energies. All in all, assuming either Kamae et al.’s or Aharonian and Atoyan’s expression for all energies does not import substantial differences, as Table 4 shows, and is computationally preferable. Finally, in Figure 13 we compare the inclusive cross sections for charged pions with experimental data up to about 1 TeV, which are again found in agreement with experimental data (except for a bunch of data points at low proton energies, in the case of positive charged pions). In any case, for situations where the density of cosmic rays or of target nuclei or both are high, neutral pion decay is expected to be the dominant process above 100 MeV, so that possible uncertainties in parameterizations of charged pions cross sections are not expected to play a significant role in the prediction of fluxes. ## Acknowledgments The work of ED-S was done under a FPI grant of the Ministry of Science and Tecnology of Spain. The work of DFT was performed under the auspices of the U.S. D.O.E. (NNSA), by the University of California Lawrence Livermore National Laboratory under contract No. W-7405-Eng-48. We thank Juan Cortina, Igor Moskalenko, and Olaf Reimer for comments. We thank the referee, V. Dogiel, for remarks that led to an improvement of this work.
warning/0506/cond-mat0506234.html
ar5iv
text
# A comparative study of high-field diamagnetic fluctuations in deoxygenated 𝑌⁢𝐵⁢𝑎₂⁢𝐶⁢𝑢₃⁢𝑂_{7-𝑥} and polycrystalline (𝐵⁢𝑖-𝑃⁢𝑏)₂⁢𝑆⁢𝑟₂⁢𝐶⁢𝑎₂⁢𝐶⁢𝑢₃⁢𝑂₁₀ ## I introduction Fluctuation effects below and above the superconducting transition temperature , $`T_c(H)`$, for high-$`T_c`$ superconductors have been the subject of intensive work . Field effects have been shown to enhance the fluctuation region (beyond Gaussian) due to the one-dimensional character imposed by a strong field to the Cooper pairs , constraining the paired quasi-particles to remain in the lowest Landau level (LLL). The corresponding quartic term correction to the free energy density suppresses the sharp second-order phase transition, rounding it on $`\mathrm{\Delta }T`$ near $`T_c(H)`$ . High- $`T_c`$ superconductors, with their small coherence lengths, $`\xi `$, high Ginzburg-Landau (GL) parameters, $`\kappa `$, and high critical temperatures, $`T_c`$, display a broad fluctuation region. Theoretical work has shown that in the presence of strong magnetic fields, for which the LLL approximation can be used, the $`\mathrm{\Delta }T`$ fluctuation region around $`T_c`$, for a system with dimensionality D is proportional to $`(TH)^{(D1)/D}`$, and that within this region physical quantities such as magnetization and direct-current conductivity should be scaled with $`(TH)^{(D1)/D}`$. For magnetization in particular, scaling predicts that $`M`$ vs $`T`$ data obtained at different fields, $`H`$, should collapse onto a single curve when the variable $`M/(TH)^{(D1)/D}`$ is plotted against $`(TT_c(H))/(TH)^{(D1)/D}`$. Here, $`T_c(H)`$ becomes a fitting parameter. This scaling law has been used to identify LLL fluctuations in a given material, and also to determine its dimensionality . An important check of the scaling is that it should provide reasonable values of $`T_c(H)`$ . For lower-dimensional or layered materials, the LLL analysis also helps to explain the crossing points observed in magnetization-vs-temperature (M-vs-T) curves . Despite the intensive work on LLL fluctuations, there have been few experimental efforts devoted to study of the universality of the high-field fluctuation scaling, which is an important prediction of the LLL-based theories for systems with a given dimensionality. Theories predicting a universal curve have been developed by Ullah and Dorsey , Tesanovic et al , and Rosenstein et al. . They have been used to fit data of specific systems. In most of the papers where a universal curve was fitted over experimental data, the fitting produced very large values of $`\frac{H_{c2}}{T}`$, values much larger (4-5 times) than experimentally determined ones. Possibly due to this fact, the literature is to our knowledge lacking on experimental work that tests adequately the existence of a universal curve that should fit two or more different samples. Furthermore, it is of fundamental interest to ask whether scaling for two different high-$`T_c`$ systems exhibits universal behavior. To provide such a test was the main objective of this work. The work was also motivated by the possibility of studying the evolution of the magnetic phase diagram of Y123 with content of oxygen. In this work, we addressed the above question by studying two different systems ($`YBa_2Cu_3O_{7x}`$ (Y123) and a highly textured polycrystalline specimen of $`(BiPb)_2Sr_2Ca_2Cu_3O_{10}`$ (Bi2223)). These two systems are expected to exhibit D = 2 for high fields. We fit a two-dimensional LLL universal curve to the experimental data from each sample, which produced reasonable values of $`\kappa `$ for three samples (except, for the Y123 sample with $`T_c`$ = 62.5 K), but values of $`\frac{H_{c2}}{T}`$ were too large when compared with the experimentally determined values. We also performed a simultaneous two-dimensional LLL scaling analysis, obtaining a single collapsed curve for all data sets. The single collapsed curve was obtained after multiplying each sample data set by an appropriate sample-dependent scaling factor. An expression of the universal two-dimensional x-axis scaling factor was extracted from the theory and compared with the experimental values. The largest deviation found from this comparison is over 40$`\%`$. The values of $`\frac{H_{c2}}{T}`$ used in the theoretical x-axis expression were the experimental ones. ## II experiment The samples that were measured were: three single crystals of $`YBa_2Cu_3O_{7x}`$ with critical temperatures, $`T_c`$ of 62.5 K (x = 0.35), 52 K (x = 0.5), and 41 K (x = 0.6), and corresponding masses of 1.2, 1.0, and 1.0 mg, and a highly textured polycrystalline $`(BiPb)_2Sr_2Ca_2Cu_3O_{10}`$ with $`T_c`$ = 108 K and m $``$ 4 mg. All of the Y123 single crystals were of approximate dimensions 1 x 1 x 0.2 mm, with the $`c`$ axis along the shorter direction, and each exhibited sharp, fully developed transitions ($`\mathrm{\Delta }T_c1K`$). The textured Bi2223 was cut into a similar shape as that of the single crystals, but with a mass about four times larger. The Bi2223 sample also exhibited a fully developed transition ($`\mathrm{\Delta }T_c3K`$). A commercial magnetometer (MPMS Quantum Design) based on a superconducting quantum interference device (SQUID) was utilized. The scan length was 3 cm, which minimized field inhomogeneities. Experiments were conducted by obtaining isofield magnetization (M) data as a function of temperature (T), producing M-vs-T curves, with values of field running from 5000 to 50000 Oe. Magnetization data were always taken after cooling the sample below $`T_c`$ in zero magnetic field. After cooling, a magnetic field was carefully applied (without overshooting), always along the $`c`$ axis direction of the samples, and M-vs-T curves were obtained by heating the sample to temperatures well above $`T_c`$, for fixed $`\mathrm{\Delta }T`$ increments. We also obtained several field-cooled curves, corresponding to cooling the sample from above $`T_c`$ to below $`T_c`$, in an applied magnetic field. This procedure allowed determination of the reversible (equilibrium) magnetization. ## III results and discussion Figure 1 shows zero-field-cooled M-vs-T data obtained for the Y123 (main figure) and Bi2223 (inset of Fig. 1) samples after background correction. Background corrections for all samples were of the type M(T) = A + B/T, where A and B are constants determined by fitting a selected region of a given curve well above $`T_c(H)`$. Two facts, observed for all curves of Y123 samples, are evident: (1) Magnetization showed a ”hump” that appeared at lower temperatures, occurring at the position at which the field-cooled and zero-field-cooled curves separated (where irreversibility set in). Such a hump appeared in all M(T) curves obtained from the Y123 samples. The hump is reminiscent of the so-called ”paramagnetic Meissner effect”,PME, observed in Type-II superconductors . In this work, magnetization was obtained through a routine that fits a standard symmetric dipole into the output signal, which was generated after the entire length of the scan was completed. We observed this output dipole signal on the screen during some measurements. The dipole diamagnetic signal for the scan length that was used was symmetric for temperatures corresponding to the reversible region of the M-vs-T curve, but the signal dramatically lost symmetry for temperatures corresponding to the hump (irreversible region). Therefore, we concluded that the hump was probably an artifact of the equipment that was used and was due to the fitting routine. It is worth mentioning that the paramagnetic Meissner effect was also observed in niobium in Ref.19 where it is suggested that the PME is a surface effect not specific to high-$`T_c`$ superconductors. In this work, the hump was used to identify the lower-temperature limit of the reversible region in each M-vs-T curve. We shall not discuss irreversibility effects or the irreversibility line in this work. (2) A rounded transition was observed as magnetization tended to zero. Such a transition impeded a straightforward determination of $`T_c(H)`$. This response was observed in all M-vs-T curves that were obtained. Such a rounded transition has been shown to be associated with high-field diamagnetic fluctuations of LLL type . Figure 1 also shows the existence of a crossing point in each data set, occurring close to $`T_c(H)`$ (clearest for the sample with $`T_c`$ = 41 K). These crossing points have been discussed previously for deoxygenated Y123 samples . In that study , the authors used measurements obtained for two samples of this work (the samples with $`T_c`$ = 62.5 K and 52 K). For the Y123 system, reduction of oxygen content decreases $`T_c`$ and increases the anisotropy $`\gamma `$ . For layered systems, the crossing magnetic field above which the dimensionality of the system is reduced is expected to decrease as anisotropy increases . Because the value of the anisotropy of Bi2223 ($`\gamma 3150`$ ), which is known to obey two-dimensional LLL scaling , is of the same order of magnitude as that of $`\gamma `$ of our deoxygenated Y123 samples , we began analysis of the Y123 data by performing a LLL scaling with D = 2. We mention that we also attempted to perform a three-dimensional version of the LLL scaling of our Y123 data. The three-dimensional version was, however, shown to apply only to a narrow region close to $`T_c(H)`$ for the samples with $`T_c`$ = 52 and 62.5 K, and it is not presented here. In the two-dimensional LLL analysis that was performed, $`M/(TH)^{1/2}`$ was plotted versus $`(TT_c(H))/(TH)^{1/2}`$. $`T_c(H)`$ became a fitting parameter, chosen to make the data collapse onto a single curve. Figure 2 shows the results of the scaling analysis performed on reversible data of Fig. 1. The values of the pairs $`(T_c(H)(K),H(10^3Oe))`$ and the correspondent value of $`\frac{H_{c2}}{T}`$ for each sample as obtained from the two-dimensional LLL analysis are listed in the table below: $$\left[\begin{array}{cccc}Y123(41K)\hfill & Y123(52K)& Y123(62K)& \hfill Bi2223\\ (41,0)\hfill & (52,0)& (62.5,0)& \hfill (108.5,0)\\ (40.5,5)\hfill & (50.8,10)& (61.5,10)& \hfill (107,10)\\ (40,10)\hfill & (48.7,20)& (60.7,20)& \hfill (105.5,20)\\ (39.3,20)\hfill & (47.3,30)& (59.4,30)& \hfill (104,30)\\ (38.2,30)\hfill & (46,40)& (58.2,40)& \hfill (102.5,40)\\ (37,40)\hfill & (44.7,50)& (57,50)& \hfill (101,50)\\ (36,50)\hfill & & & \\ \frac{H_{c2}}{T}\hfill & \frac{H_{c2}}{T}& \frac{H_{c2}}{T}& \hfill \frac{H_{c2}}{T}\\ 10kOe/K\hfill & 7kOe/K& 9kOe/K& \hfill 7.6kOe/K\end{array}\right]$$ It is worth mentioning that these values of $`T_c(H)`$ are in reasonable agreement with values of $`T_c(H)`$ estimated for each M-vs-T curve from the standard linear extrapolation of the reversible magnetization down to M = 0 . The obtained value of $`\frac{H_{c2}}{T}`$ for Bi2223 is in good agreement with published values . The results of the two-dimensional LLL analysis presented in Fig. 2 show that the collapse of data was better for fields in the region of 20000-50000 Oe for the crystal with $`T_c`$ = 62.5 K, for fields in the region 10000-50000 Oe for the crystal with $`T_c`$ = 52 K, and for fields in the region 5000-50000 Oe for crystal with $`T_c`$ = 41 K. It is interesting to note that the lowest-field data obeying the two-dimensional LLL for Y123 decreased as the anisotropy of the system increased. This behavior is predicted by the theory. For layered systems, the crossing magnetic field above which the dimensionality of the system is reduced is expected to decrease as anisotropy increases . The literature presents an analytical expression for the two-dimensional LLL scaling function of the magnetization, which is given by Eq. 7 in Ref. 4 : $$\frac{M(H,T)}{(HT)^{1/2}}\frac{s\varphi _0}{A}|\frac{H_{c2}}{T}|=x\sqrt{x^2+2},$$ where $`x=A(TT_c(H))/(TH)^{1/2}`$, $`A=a^{}(\varphi _0s/2b)^{1/2}U_0`$, $`\varphi _0`$ is the quantum flux, $`s`$ is the interlayer distance, $`U_00.8`$ (around $`H_{c2}(T)`$), and $`a^{}(TT_c)=a(T)`$, where $`a`$ and $`b`$ are the GL coefficients. We applied this expression to our data. Fits by a least-squares minimization method were obtained with magnetization in Gauss units. Dashed lines in Fig. 2 represent the fitting results on the respective data. We used $`d_{Y123}=6.38g/cm^3`$ and $`s_{Y123}=12\AA `$. To our knowledge, the literature presents only one similar study (two-dimensional LLL fitting) performed in deoxygenated Y123, with $`T_c50K`$ . Bi2223 has been already studied and reported on in Ref.5. The fitting parameters are $`(\frac{H_{c2}}{T})_{T=Tc}`$ and $`a^{}/(2b)^{1/2}=\frac{1}{(\sqrt{8}\pi }(\frac{H_c}{T})_{T=Tc}=\frac{1}{4\sqrt{\pi }\kappa }(\frac{H_{c2}}{T})_{T=Tc}`$, where we used the relation $`H_{c2}=\sqrt{2}\kappa H_c`$ obtained in the GL linear approximation , where $`H_c`$ is the GL critical field. The fitting results are shown in Fig. 2 and produced the values listed below: $$\left[\begin{array}{cccc}Y123(41K)\hfill & Y123(52K)& Y123(62K)& \hfill Bi2223\\ \kappa =79\hfill & \kappa =73& \kappa =114& \hfill \kappa =96\\ |\frac{H_{c2}}{T}|\hfill & |\frac{H_{c2}}{T}|& |\frac{H_{c2}}{T}|& \hfill |\frac{H_{c2}}{T}|\\ 21652Oe/K\hfill & 19875Oe/K& 3858Oe/K& \hfill 23224Oe/K\end{array}\right]$$ It is important to note two interesting results of the fittings. (1) It is consistently suggested that as the system becomes more two dimensional (as $`T_c`$ of Y123 drops), fitting produces a value of the parameter $`\kappa `$ that is closer to the experimentally estimated value. This fact is, from our knowledge, a new result. (2) Despite the good quality of the fittings, the experimental values of the superconducting diagmagnetic fluctuations above $`T_c(H)`$ are smaller than the theoretical ones. These facts were observed for data of Y123 samples with $`T_c`$ = 52 and 41 K and for Bi2223. The possibility that the theory overestimates thermal fluctuations on the magnetization of high-$`T_c`$ superconductors, as suggested by the data analysis, has been discussed in Ref. 28. Although the fittings produced curves which reproduced quite well the experimental data (see the dashed lines in Fig. 2), with reasonable values of $`\kappa `$ (with the exception for Y123 with $`T_c`$= 62.5 K), the resulting values of $`(\frac{H_{c2}}{T})_{T=T_c}`$ are 2-3 times larger than the corresponding experimental values that we determined. Because of this discrepancy, we then used a different approach to interrogate the predicted universality of the two-dimensional LLL scaling. We present in Fig. 3 a simultaneous scaling curve of all samples data sets, including reversible data obtained for the Bi2223 textured sample. The resulting collapsed curve was obtained after multiplying the collapsed curve obtained from the two-dimensional LLL analysis of each sample, as shown in Fig.2, by an appropriate sample-dependent scaling factor. Specifically, the $`x`$ axis and $`y`$ axis of Y123 and Bi2223 data were multiplied by the factors: x(0.84) and y(2.04) for Y123 with $`T_c`$ = 62.5 K, x(0.70) and y(1.61) for Y123 with $`T_c`$ = 52 K, x(0.67) and y(1.79) for Y123 with $`T_c`$ = 41 K, and x(1.18) and y(0.69) for Bi2223. The collapse of data from the four different samples, belonging to two different systems, was good and at a first interpretation, may provide support to the notion of universality. The dashed lines in Fig. 3 (which appear as a single fitting curve in Fig. 3 because they are too close to be resolved individually) represent the fittings already presented in Fig. 2. We can take a step further, and analyze the collapse shown in Fig. 3 by comparing the scaling variables of each axis. Comparison of the ”magnetization-axis” ($`y`$ axis) is not directly possible because of differences in sample geometry. For example, small differences in thickness between samples, as was the case for these measurements, give rise to different demagnetization factors. In any case, one may observe that the scaling factors of the $`y`$ axis were roughly proportional to sample mass ratio. Comparison of the scaling variable along the $`x`$ axis is possible, because Ullah and Dorsey provide an analytical expression for the full two-dimensional LLL temperature scaling variable along the $`x`$ axis: $$x=\frac{s}{(2\kappa ^21)^{1/2}}|\frac{H_{c2}}{T}|\frac{(TT_c(H))}{(TH)^{1/2}},$$ where s is the interlayer spacing and $`\kappa `$ is the Ginzburg-Landau parameter. The ratio of the respective Bi2223 and Y123 scaling factors, $$r=(\frac{s}{(2\kappa ^21)^{1/2}}|\frac{H_{c2}}{T}|)_{Bi2223}/(\frac{s}{(2\kappa ^21)^{1/2}}|\frac{H_{c2}}{T}|)_{Y123}$$ can be used to rescale the Bi2223 data to make them collapse onto the Y123 data. The experimental values of $`r=x_{Bi}/x_{Y123}`$, where $`x_{Bi}`$ and $`x_{Y123}`$ are the $`x`$ axis sample-dependent factors listed above. A comparison of both $`r`$ values for the samples that were examined can test the two-dimensional LLL hypothesis of universality. For a direct comparison with the experimental values, we require values of $`\kappa `$, which for Y123 can be estimated from isothermal $`MvsH`$ curves previously obtained for our Y123 samples . The inset of Fig. 3 shows selected isothermal $`MvsH`$ curves. Measurements of additional $`MvsH`$ curves in these samples could not be made reliably because of the current low-field restrictions of our equipment. The curves in the inset of Fig. 3 allow one to estimate values of $`H_{c1}`$. This value was obtained as the field for which the magnetization curve deviates from the linear Meissner region. These values, coupled with the GL relation, $`H_{c1}/H_{c2}=\mathrm{ln}\kappa /2\kappa ^2`$ and values of $`H_{c2}`$ obtained in this work from the 2D-LLL scaling analysis (values are listed above), then allow one to estimate values of $`\kappa `$. For the sample with $`T_c`$ = 62.5 K, we also estimate values of $`H_{c1}`$ from $`MvsH`$ curves obtained at low temperatures , which allow us to obtain the extrapolated value of $`H_{c1}(T=0)`$. The estimated values of $`\kappa `$ are: $`\kappa 65`$ for Y123($`T_c`$=52 K), $`\kappa 56`$ for Y123($`T_c`$=62.5 K) and $`\kappa 80`$ for Y123($`T_c`$ = 41 K). The error on $`\kappa `$ is (over) estimated to be $`10`$. By using the experimental obtained values of $`\frac{H_{c2}}{T}`$, $`\kappa `$ for Y123, $`s_{Y123}`$=$`12\AA `$, $`\kappa _{Bi}`$ = 82 , and $`s_{Bi}`$=$`20\AA `$ in the expression above for $`r`$, we obtain: $`r(\frac{Bi}{Y62K})`$= 0.96 , $`r(\frac{Bi}{Y52K})`$= 1.43 and $`r(\frac{Bi}{Y41K})`$ = 1.23. The experimental values are $`r(\frac{Bi}{Y62K})`$ = 1.40, $`r(\frac{Bi}{Y52K})`$ = 1.69, and $`r(\frac{Bi}{Y41K})`$ = 1.76. The largest deviations between experimental and calculated values are $`40\%`$, which is found for samples with $`T_c`$= 62.5 and 41 K. To provide conclusive evidence of two-dimensional LLL scaling universality, one would expect much smaller deviations in $`r`$. It should be noted that, independently of the above result, the basic scaling theory is expected to be independently valid for each system here studied. We finally note that in previous work with Nb , where simultaneous scaling was performed for Nb and Y123 data (two systems with dimension = 3), evidence for universal three-dimensional LLL behavior was not obtained. Finaly, we attempt to verify if diamagnetic fluctuations due to non-local electrodynamics effects , that appear from strong magnetic fields, might explain the observed deviations between theory and experiment concerning two-dimensional LLL. It is important to mention that the LLL theory was developed under the Ginzburg-Landau theory for strong magnetic fields which is local and corresponds to the first term expansion of the Gorkov theory. For stronger fields , higher order terms expansion of the Gorkov theory are necessary producing non-local field terms. In this case, it has been shown that $`M(T)H^{0.5}/T`$ is a universal function of $`H/H_s`$, where $`M(T)`$ is the fluctuation magnetization and $`H_s`$ is a sample dependent field. Experimentally, the above scaling only works at $`T=T_c`$. Above $`T_c`$, $`M(T)H^{0.5}/T`$ vs $`H/H_s`$ produced universal curves at temperatures given by $`dH_{c2}dT(TT_c)/H=c`$ where $`c0`$ is a constant. This universal behavior have been shown to work for many convencional type II superconductors with $`H_s`$ as high as $`H_{c2}`$, but, from our knowledge, it was never tested on high field data of high-Tc superconductors. Figure 4a shows curves $`M(T)H^{0.5}/T`$ vs $`H/H_s`$ obtained for Y123 data shown in Fig. 1. We only analysed Y123 data because they have a sharper $`T_c`$ when compared to Bi2223 (in the analysis of convencional superconductor data, $`\mathrm{\Delta }T_c10^3K`$ ). It is possible to see in Fig. 4a that the behavior of the curve obtained with values of magnetization at $`T=T_c`$ is followed by the curves with $`dH_{c2}dT(TT_c)/H=0,5`$ and $`1`$, which resembles the behavior observed in the similar plot for convencional superconductors . It is not possible to verify if our data follow the same universal curve of Ref.32 since the magnitude of $`M(T)H^{0.5}/T`$ for our samples are two orders larger than the same quantity for conventional superconductors. In plotting Fig. 4a, we try to obtain values of the X axis with similar order of magnitude (aproximately one order larger here) as in the universal curve obtained for convencional superconductors appearing in Ref.32, which produced the values of the field $`H_s`$ listed in Fig. 4a. It is interesting to note that the data for Y123 with $`T_c=62K`$ do not follow the ”scale” for the curves with $`dH_{c2}dT(TT_c)/H=0,5`$ and $`1`$. This is because data for this sample ($`T_c=62K`$) have a (almost) perfect match with the 2D-LLL theory (see Fig. 2). At $`T=T_c`$, we observed that the experimental values of magnetization are close to the values calculated from the 2D-LLL expression, which means that only at $`T=T_c`$ the fluctuation magnetization obtained from the 2D-LLL theory also satisfies the non-local field scaling of Fig. 4a. The later fact does not occurr for temperatures above $`T_c`$. Figure 4b shows a plot for samples with $`T_c`$=52 and 41 K (similar to Fig. 4a) for $`dH_{c2}dT(TT_c)/H=1`$ , but with magnetization calculated from the 2D-LLL expression using the fitting results ($`dH_{c2}dT`$ and $`\kappa `$) listed in the table (the curves calculated for $`dH_{c2}dT(TT_c)/H=0.5`$ are very similar and are not shown). It is clear in Fig. 4b that, for temperatures above $`T_c`$, fluctuation magnetization calculated from 2D-LLL for samples with $`T_c`$=52 and 41 K do not follow the non-local field scaling that experimental data do in Fig. 4a. Comparison of Figs. 4a and 4b suggests that non-local electrodynamics effects due to high fields should have an important rule on fluctuation diamagnetism above $`T_c`$ for these two samples, which might explain the deviations between the experiment and lowest-Landau-level theory found in this work. In conclusion, our study shown that high field fluctuations on reversible data of Y123 crystals and Bi2223 are well explained by two-dimensional lowest-Landau-level fluctuations. The scaling analysis performed on Y123 data produced consistent values for the upper critical field $`H_{c2}`$ for the studied samples. The data of each sample was fitted by the predicted 2D-LLL universal theory, which produced reasonable values of the Ginzburg-Landau parameter $`\kappa `$, but much larger values of $`\frac{H_{c2}}{T}`$. The fitting show that theory predicts larger values for the diamagnetic fluctuation for temperatures above $`T_c(H)`$. A different test, than the fitting, was performed to check the hypothesis of universality of 2D-LLL, but the results did not support this hypothesis for the studied samples. We also observe that Y123 magnetization data for temperatures above $`T_c`$ obbey a universal scaling obtained for the diamagnetic fluctuation magnetization from a theory considering non-local field effects. The same scaling was not obbeyed by the corresponding magnetization calculated from the two-dimensional lowest-Landau-level theory. We grateful thank the Referee who suggested us to verify the possible importance of non-local eletrodynamics effects on the fluctuation diamagnetism in our data. We also thank Mark Friesen for helpful discussions. This work was partially supported by CNPq, Brazilian Agency. The work at Argonne National Laboratory was supported by the U.S. Department of Energy, under Contract W-31-109-Eng-38. $``$ Corresponding author. E-mail: said@if.ufrj.br Figure Captions Fig1. Zero-field-cooled M-vs-T curves as obtained for $`YBa_2Cu_3O_{7x}`$ crystals for magnetic fields ranging from 5000 to 50000 Oe for crystal with $`T_c`$ = 41 K, and from 10000 to 50000 Oe for crystals with $`T_c`$ = 52 K and 62.5 K. The inset shows zero-field-cooled magnetization data of Bi2223 used in the scaling analysis, obtained for magnetic fields ranging from 10000 do 50000 Oe. Fig2. Two-dimensional LLL scaling of Y123 and Bi2223 (inset) data presented in Fig. 1. Dashed lines represent fittings of the data to the two-dimensional LLL universal expression for magnetization. Y123 curves were displaced along the Y axis to appear in the same figure. Fig3. Simultaneous two-dimensional LLL scaling of Y123 and Bi2223 (inset) data presented in Fig. 1. Dashed lines represent fittings of the data to the two-dimensional LLL universal expression for magnetization. The inset shows selected isothermic zero-field-cooled magnetization curves for the Y123 crystals. The magnetization curves were displaced along the Y axis to appear in the same figure, and magnetization for Y123 with $`T_c`$= 41 K was multiplied by 2. Fig.4 (a)$`M(T)H^{0.5}/T`$ is plotted against $`H/H_s`$ for Y123 data of Fig. 1. Dashed lines are only a guide to the eyes. (b)$`M(T)H^{0.5}/T`$ is plotted against $`H/H_s`$ with values of M(T) obtained from the two-dimensional lowest-Landau-level expression.
warning/0506/astro-ph0506761.html
ar5iv
text
# Stephan’s Quintet with XMM-Newton ## 1 Introduction Multi-wavelength data continue to accumulate for Stephan’s Quintet (SQ) the most studied of the compact groups of galaxies. A framework for these observations has developed that may be quite generally relevant to the compact group phenomenon. Groups like SQ represent the densest galaxy aggregates in the non-clustered Universe. If SQ is representative, then compact groups evolve and grow through a process of sequential harassment by, and acquisition of, infalling intruder galaxies. It is not clear what stimulates the infall because the observed baryonic mass in SQ is about an order of magnitude too small to account for it. Two intruder galaxies have been identified in SQ (Moles et al. 1998). As summarized in Trinchieri et al. (2003), the older, and likely captured, intruder (NGC 7320c) has left evidence of two passages in the form of parallel tidal tails. NGC 7319 was most likely the victim of the last passage having lost most of its ISM. From the evidence of an “UV loop” in the new GALEX images of this system, Xu et al. (2005) suggest that only the older tail results from the passage of NGC 7320c, the younger one being caused by a close encounter with NGC 7318a. In either case, the Seyfert 2 nucleus in NGC 7319 may represent another manifestation of the encounter. Both it and NGC 7320c are examples of galaxies in transition from later (S) to earlier (E/S0) type. One or both of the members of SQ with elliptical morphology (NGC 7317 or 7318a) may have been stripped in earlier passages of the old intruder. The “new” intruder (NGC 7318b) recently penetrated SQ at unusually high velocity ($`\mathrm{\Delta }`$V$``$1000km/s) which suggests that it is not bound to the system. The strongest signs of current activity are related to this latter encounter. The results of the active dynamical history in SQ yield multifold observational manifestations of the interaction process: 1) tidal deformations, 2) ISM stripping, 3) star formation suppression within the galaxies, 4) star formation sites in the stripped debris, 5) morphological transformations, 6) creation of a common halo of stripped stars, 7) shocks and, possibly, AGN ignition. It is widely assumed that interactions were more common in the early Universe and this makes detailed study of groups like SQ important beyond their benefit as laboratories for studying interaction physics. The signatures of the above interaction events have now been observed in all windows of the electromagnetic spectrum except the gamma ray. SQ might therefore become a reference for extragalactic observers. Recent Chandra observations of SQ (Trinchieri et al. 2003) generally support the above evolutionary picture and have shown quite distinctly the effects of interaction between SQ and its intruders. A large scale shock ($``$40kpc at 85 Mpc) seen in both ROSAT-HRI and Chandra images as a prominent narrow NS feature embedded in more complex and extended diffuse emission is further confirmed by these XMM-Newton observations (Fig. 1). The NS structure was found to be distinctly clumpy and sharply bounded on the W side. Spatial coincidence of the X-ray shock is found with both radio continuum (Williams et al. 2002) and forbidden \[N ii \] emission (Sulentic et al. 2001) while little correspondence is seen between the X-ray and optical broad band images. The principal X-ray features are best explained as manifestations of the shock produced by collision of the new intruder with the debris field produced by passages of the old intruder. However, if the stripped ISM of NGC 7319 was largely displaced towards the east, in the direction of NGC 7320c, as suggested by the observed H i distribution (Williams et al. 2002), then currently shocked debris field must be the product of an earlier encounter between the old intruder and, perhaps, NGC 7318a. The NS X-ray feature was interpreted (Trinchieri et al. 2003) as a bow shock propagating through the pre-existing debris field and heating it to a temperature of $`0.5`$ keV. The low temperature of the post-shock is a problem unless we postulate: a) an oblique shock and/or b) a weak shock where the upstream medium is hot and has a sizable counter-pressure. As discussed in Trinchieri et al. (2003), a standard weak shock is unlikely given the high Mach number of the inflowing gas in the frame of reference of the new intruder. In order to reduce the post-shock temperature to a value consistent with that derived from X-ray spectral fitting, the gas would have to enter the shock with respect to the shock surface at an angle of $`30^{}`$ or less. This will only happen at the wings of the bow-shock. In Trinchieri et al. (2003) we estimated the cooling time from X-ray spectral fitting and a standard cooling function for a low metallicity plasma in collisional ionization equilibrium to be roughly $`4\times 10^8`$ yr for a mean density of $`n_X2.7\times 10^2\mathrm{cm}^3`$. Note that in the surrounding upstream medium the density can be lower by a factor of $`510`$, so that cooling can be completely negligible, but that denser regions in the shock have higher densities and correspondingly shorter cooling times. This interpretation is attractive because it would also better explain evidence that the new intruder shows considerable tidal distortion. In this case in flagrante delicto implies an intruder crossing time of t$`{}_{c}{}^{}10^8`$ years. ## 2 XMM-Newton Data Analysis SQ was observed for $``$40 ks with EPIC-MOS and EPIC-pn during revolution 366 of XMM-Newton. The event files were reprocessed with the $`xmmsas\mathrm{5.4.1}`$ release and further cleaned of residual high background due to flaring activity. We used the EPIC-pn light curve for processing data from all instruments. Only photons with FLAG=0 and PATTERN$``$4 (for EPIC-pn), or PATTERN$``$12 (EPIC-MOS), were used in the analysis. The final exposure times for the accepted events are listed in Table 1. Data analysis employed both the $`xmmsas`$/ftools and $`ciao`$/funtools tasks with heavy reliance on ds9 and corollary software. ### 2.1 X-ray Images: characterization of the diffuse emission In an effort to enhance the signal-to-noise we merged the central CCD frames of the two EPIC-MOS observations and smoothed the result with an adaptive algorithm ($`ciaocsmooth`$). EPIC-pn data are kept separate because of the different patterns of CCD gaps in EPIC-pn and EPIC-MOS instruments. EPIC-pn, EPIC-MOS and Chandra (Trinchieri et al. 2003) images (0.3-3.0 keV) are compared in Fig. 1, which illustrates the complimentary power of data from the two satellites. While the Chandra image clearly shows more discrete features, the large collecting area of XMM-Newton facilitates detection of diffuse X-ray emission components. A comparison of EPIC-MOS emission in different energy bands is shown in Fig. 2. Both figures reveal the complexity of the emission in SQ with a multitude of both compact and extended features centered on the NS oriented shock, with different relative strengths as a function of the energy band considered (see later). A previously undetected component is visible towards the SE as an extension that connects the low surface brightness emission at the center of the group with an enhancement that was marginally detected in the Chandra data (Trinchieri et al. 2003). This new component, called TAIL for easy reference in the text, is faint even in XMM-Newton images making a full characterization of its properties very difficult. The more sensitive XMM-Newton data allow a better definition of low surface brightness emission in SQ and we have tried to determine its extent by means of radial profiles of the X-ray emission. Regions of CCD gaps, bad columns and bad pixels identified on the exposure maps are masked and all detected sources outside a radius of $`3^{}`$ from the center of SQ are excluded. Emission above $`2.53.0`$ keV is associated with the nucleus of NGC 7319 and a few much fainter individual sources. Diffuse emission is detected out to a maximum extent of r$`4^{}`$ at energies below $`2.5`$ keV. However, the emission becomes clearly nonuniform with increasing distance from the central shock region, so that the extent depends on the assumed direction. Figure 3 shows both the clear excess in the TAIL region identified above (left panel) and the azimuthal dependence of the extent of the emission (right panel), towards the SE (TAIL) and the SW (towards NGC 7317) at an average radius of $`2^{}3^{}`$. The smaller extent towards S could reflect the presence of additional absorption, due in part also to the foreground NGC 7320 (however, the same should be true for TAIL). The EPIC-MOS CCD gaps make any characterization of any diffuse structure outside r$`4^{}`$ difficult to quantify properly. In order to better highlight different spectral signatures in the complex emission from SQ we have generated a color image combining data in three broad bands: 0.3-1.5, 1.5-2.5 and 2.5-7.0 keV (displayed as red, green, and blue respectively). The resulting image, shown in Fig. 4 shows a number of localized regions with different colors that imply differences in photon energy distribution. Many of these regions lack sufficient photons for a detailed spectral analysis but we can make a rough comparison of the relative photon distributions for some of the regions that are illustrated in Fig. 5. A more detailed analysis is given for regions with high enough photon statistics (see § 2.2). Constant broad energy bins are used for constructing all of the “pseudo-spectra” displayed in Fig. 6. These energy distributions have not been corrected for instrumental response. They were extracted from relatively small areas of the central CCD in EPIC-MOS so we do not expect variations in the response matrix to affect this qualitative comparison between different sources and regions, but cannot be used to derive spectral shapes. The pseudo-spectra show significant differences that suggest these sources are unlikely to be fit with the same spectral model. We comment on the most striking: 1) Regions B and C (top panel) show similar distributions especially compared with region A. The average pseudo-spectrum of the shock region is intermediate between A and B distributions. Chandra data suggest that a compact source (Chandra #5 Trinchieri et al. 2003) is embedded in condensation C (but not at its center). With the current astrometry and published positions of the H$`\alpha `$ knots (e.g. Xu et al. 2003), the Chandra position does not suggest coincidence with any obvious optical feature in the new intruder, so it might be an unrelated background source. A region of intense star formation (STARBURST A in Xu et al. 1999) and associated H i and CO emission lie between the N end of the shock and source C. 2) The sources in the middle panel are associated with SQ galaxies and show similar pseudo-spectra. NGC 7319 shows a strong hard photon excess which is likely related to an obscured active nucleus. Significantly different line-of-sight absorbing column densities might explain the large difference between the lowest energy points of NGC 7318a and NGC 7317. This is consistent with the lack of any radio line or continuum detections near to NGC 7317. Unresolved radio continuum emission is detected from NGC 7318a (Williams et al. 2002; Xu et al. 2003). Source # 2 is also included in this panel. Thus is the only one of four very large (D$``$400 pc) H$`\alpha `$ emission condensations detected as a discrete X-ray source (Trinchieri et al. 2003) and belongs to NGC 7318b (Sulentic et al. 2001). 3) Diffuse emission regions outside the central shock region (bottom panel) do not show identical spectra: TAIL and D are consistent with each other but not with HALO. The latter feature is more reminiscent of the spectral distribution of the shock front and may be evidence for a large scale signature of the shocked new intruder disk. This will complicate any inferences about the extent and integrated properties of diffuse X-ray component that existed before the arrival of the new intruder. It is unclear if D is a condensation near the end of this tail or an unrelated background source: there is evidence of an extended object at its center (redshift and nature unknown at the present time), but at the same location there is a clear deficit in the H i distribution that, if foreground, would represent a smaller absorbing column with consequent local enhancement of the emission. The location of D also coincides with the “end” of the H i in Arc-S (Williams et al. 2002) and could also represent the edge of the expansion of the hot gas. The similarity of the spectral distributions for D and TAIL might favor the interpretation of D as a feature associated to the more diffuse component in TAIL. ### 2.2 Characterization of individual features The previous images and pseudo-spectra enable us to identify emission regions where it is possible to derive both the contribution of individual components (see Table 2) and where the photon energy distribution allows a more detailed spectral analysis. Table 2 lists detected counts, fluxes and luminosities for individual sources detected and already presented in Fig. 5 for which a more detailed spectral analysis is not possible. Regions with stronger emission are instead considered for more detailed spectral modeling, as summarized in Table 3. Source shapes and sizes were adjusted to avoid regions of non-uniform illumination (e.g. CCD gaps), but were kept as large as possible (compatible with colors and morphologies implying coherent components, as suggested by both our X-ray color image and Fig. 6) in order to improve the signal-to-noise in the derived spectra. Photons were binned to achieve a minimum signal-to-noise $`23`$ (depending on the resulting number of bins) in each energy bin after background subtraction. We generated $`arf`$ and $`rmf`$ files, using the most recent calibrations. We used PATTERN$``$12 for EPIC-MOS\[1-2\] (singles, doubles, triples and quadruples) and PATTERN=0 for EPIC-pn (singles only), since we are mainly interested in softer energies where the “single” percentage was highest. We also considered “doubles” in the Seyfert data that contains many photons at higher energies in order to increase the signal-to-noise. Results of spectral modeling are summarized in Table 3. The data do not provide meaningful constraints on the choice of spectral model, with several different assumptions yielding equally good results. In particular, we have made two different assumptions for the abundance parameter, fixing it at 30% of cosmic (consistent with the choice in Trinchieri et al. 2003), and at 100%. The former value gives in general better fits, i.e., lower $`\chi ^2`$ values: an additional component is needed to obtain a similar $`\chi ^2`$ under the assumption of 100% cosmic; we have chosen a power law with fixed slope at $`\mathrm{\Gamma }=1.7`$, but other choices give equal results. The spectral parameters are generally consistent in the two models. In fact the error regions are ill-defined in several cases, due to apparent non-monotonicity in the $`\chi ^2`$ space. As discussed later (see also Trinchieri et al. 2003), the spectral parameters should be regarded as only indicative, as all regions are probably complex, and might contain different components at different temperatures. Moreover, plasma temperatures are derived from models of optically thin emission, in which the plasma is assumed in collisional ionization equilibrium. In the presence of non-equilibrium plasma, such as in cases of shocks, these models would have the effect of “artificially” requiring a multi-temperature fit. The background is chosen from a region to the E in the adjacent chip for EPIC-pn (but within the same CCD for EPIC-MOS) where we see and expect no emission from SQ. We have also considered other regions to the N and W (also in adjacent EPIC-pn CCDs) but found no significant variations. Table 3 also lists the luminosities of the different regions for the different best fit parameters. In a few cases, where the best fit absorption is different, the softer band luminosity could be significantly different in the 30% and 100% abundance assumption. SHOCK region. The shock region is a clumpy NS-oriented elongated feature (Fig. 5). This structure is similar to the one observed with Chandra. It includes condensations A and B but not C. The new X-ray data require a two temperature plasma, at $`0.3`$ and $`1`$ keV with N<sub>H</sub> relatively high (3$`\times `$ the line-of-sight value). Modeling the whole region is of limited significance given the differences observed in the photon spectral distributions of the sub-regions highlighted by the color image and the pseudo-spectra (Fig. 4-6). Analysis of different condensations within the shock confirms their different spectral properties. We have considered three separate regions as indicated in Table 3 corresponding to the three main peaks determined by Chandra and confirmed in XMM-Newton. The middle one is larger than, but inclusive of, A. Assuming again plasma models, the middle and southern blob can be modeled with a single temperature but with different values for kT (0.6 and 0.2 keV respectively) and significantly different line-of-sight N<sub>H</sub>. This may reflect in part the larger H i column ($`>`$6$`\times `$10<sup>20</sup> atoms cm<sup>-2</sup>; Williams et al. 2002) along the line of sight due to the foreground galaxy NGC 7320 that partially covers the southern tip of the shock region. A hard excess is visible in the data for region A (Fig. 7), regardless of the abundance parameter, that can be accounted for by a second component (any model will do), even though the fit does not formally require a two component model ($`\chi _{min}^21`$). No equivalent component is required in the other regions (except for the 100% abundance assumption, but see above). A two temperature model, with a relatively high amount of absorbing material, is instead required for the northern tip. HALO region. All high surface brightness regions related to the shock, as well as discrete sources identified in Fig. 5, were excluded from the photon distribution yielding our best attempt to isolate a pure halo component. A two temperature model could represent the data, as in the previous region, but with a marginally credible $`\chi ^2`$=262 for 198 Degrees of Freedom, and high temperature values of kT<sub>1</sub>=0.6 and kT$`{}_{2}{}^{}=4.2`$ keV, with lower than galactic absorption N<sub>H</sub> = 0.04 (for 30% cosmic abundance; at 100% cosmic, $`\chi ^2`$=274, for similar spectral parameters). Addition of a third component brings a significant improvement in the fit quality and the temperature values closer to those observed in other regions (see Table 3). In particular, the addition of a power law component gives a $`\mathrm{\Delta }\chi ^2`$ of 45-55 (30-100% abundance respectively) for 2 additional parameters. There is a likely possibility of contamination from discrete sources (one such was detected by Chandra, # 4 in Trinchieri et al. 2003, but not by XMM-Newton). However contamination should affect all regions equally, and we have subtracted all visible sources, so any contamination should be of low level. TAIL region. Counts from both source D and the connecting region were considered in order to improve the signal-to-noise. This combination is supported by the pseudo-spectra plot (Fig. 6) that indicates a similar photon distribution for the two features. On the other hand it could be dangerous because we do not know if D belongs to SQ. Even taking this risk we still have a small number of photons and the resulting parameters are not well determined. Although a single temperature model could be marginally acceptable ($`\chi _{min}^2`$=1.4 for 34 DoF), the best fit parameters (N<sub>H</sub> significantly smaller than the line-of-sight value, kT$`>2`$ keV) are probably unreasonable for the system. A two component model gives a significantly better fit ($`\mathrm{\Delta }\chi ^2>`$ 14 for 2 additional parameters yielding an $`ftest`$ probability P$`>99.99`$ that the additional model component give an improved fit). Table 3 presents the results for a 2-plasma model fit. Equivalent results are found with plasma+thermal or plasma+power law models. NGC 7319. The spectrum of NGC 7319 is clearly complex and requires several components as we already found in our Chandra analysis. We now know that the discrete Chandra source #9 located 8 arcsec S of the nucleus is a z=2.2 QSO (Galianni et al. 2004) however it contributes only about 5% of the photons near to NGC 7319. The Seyfert nucleus was fit with a double power-law and appropriately red-shifted FeK$`\alpha `$ line (see also Trinchieri et al. 2003). The slope of the power law can in principle be better constrained now because of the higher signal-to-noise at high energy. We find however that the complex spectrum requiring a model with several independent components and large absorbing column density gives too small leverage for an accurate determination of the continuum. We also note that the region around the FeK line is not well modeled with the possibility of additional line components. At lower energies the data require a thermal plasma model with T 0.6 keV and no absorption above the galactic value, in agreement with the more tentative Chandra results. The Seyfert nucleus dominates the emission at all energies. At soft energies (0.2-2 keV) the unabsorbed power law and plasma components contribute equally while above 2 keV only the nuclear power-law source is present. ## 3 Discussion Owing to the large effective area of the telescope, the XMM-Newton observations reveal new faint X-ray features in SQ that have hitherto been missed with Chandra. In addition, improved photon statistics allows the determination of spectral components to higher accuracy, although we still cannot properly understand the real physical state of the gas, most likely because it is highly inhomogeneous and possibly not in collisional equilibrium. XMM-Newton observations support most of the conclusions based on the Chandra observations and add confidence to some of the more tentative suggestions that were proposed. The complexity of the X-ray morphology is reinforced by similarly complex spectral characteristics. Nearly all Chandra sources are detected again by XMM-Newton (allowing for the degraded spatial resolution) and we can identify additional interesting features. Each of the principal emission features in the central region of SQ will be briefly discussed here with an emphasis on new issues raised by the XMM-Newton observations. ### 3.1 X-ray Emission from SQ Member Galaxies The central regions of most SQ members are now X-ray detected with luminosities above 10<sup>39</sup> erg s<sup>-1</sup>. The exceptions involve the old (NGC 7320c) and new (NGC 7318b) intruders. NGC 7320c is far from the current center of activity and shows no detectable H i or X-ray emitting gas. The case of NGC 7318b, at the site of current impact, is more complex, and might be strongly related to the SHOCK itself (see discussion in the next section). Remnants of the H i disk of NGC 7318b are still recognizable, but no X-ray emission clearly associated with the central regions of this galaxy is detected. Only one of four bright emission line regions belonging to its southern spiral arm (Sulentic et al. 2001) is detected (source #2), at a luminosity of $`7\times 10^{39}`$ erg s<sup>-1</sup>. While high, examples of luminosities such as this are now found in increasing number associated with star forming regions (Zezas et al. 1999; Roberts & Warwick 2000; Fabbiano et al. 2001; Zezas & Fabbiano 2002; Wolter & Trinchieri 2004). Several models suggest that interaction/collisions could enhance the star formation activity in galaxies (e.g. Jog & Solomon 1992; Fujita & Nagashima 1999; Bekki & Couch 2003), so the real question becomes what distinguishes this from the other bright regions nearby: perhaps this is simply denser/more compact or the first to evolve. Both intruders, like NGC 7319 before them, are likely undergoing transformation from late-type spiral to early-type E/S0 morphologies. Perhaps NGC 7317 and 7318a underwent a similar process in past epochs. The older optical tidal tail with apparently associated H i may be a manifestation of that past activity. Given its “field” (i.e. low galaxy surface density) environment, SQ is rich in early-type galaxies and will soon grow even richer. This appears to be one of the characteristics of compact groups at least as defined by the Hickson (1982) catalog. The stripping events creating early type galaxies has given rise to large quantities of cold (Williams et al. 2002) and hot diffuse gas. The stripping events involving the intruders were apparently efficient enough to prevent the fueling of any nuclear activity so far. NGC 7319 is the exception and shows a “typical” Seyfert 2 nuclear spectrum that can be modeled by a double power-law with $`\mathrm{\Gamma }1.3`$ plus a heavily absorbed component. The derived power-law slope is flatter than inferred from the Chandra data but consistent with spectra of other Seyfert 2s (della Ceca et al. 1999; Moran et al. 2001). A thermal component is required to account for excess at low energies. This was also suggested by the Chandra data, in particular when the contribution from a circumnuclear region of more extended emission is included in the spectrum. XMM-Newton resolution means that any nuclear spectrum includes the circumnuclear region (and a negligible one from the high z quasar, see Galianni et al. 2004). ### 3.2 The Shock The sharp NS feature between NGC 7318ab and NGC 7319 shows a complex morphology with equally intriguing spectral characteristics. Higher resolution Chandra data showed several condensations and a steep W edge to the emission. These features are confirmed by XMM-Newton albeit with lower resolution. Comparison with data at other wavelengths illustrates the complexity of the shock region with shock-related effects prominent at radio (van der Hulst & Rots 1981; Williams et al. 2002; Xu et al. 2003), X ray (Pietsch et al. 1997; Sulentic et al. 2001; Trinchieri et al. 2003) and H$`\alpha `$+\[N ii \]. Some of the H$`\alpha `$+\[N ii \] emission shows a star formation signature (both at SQ and new intruder velocities) which is now also seen in the UV (Xu et al. 2005). Examination of the high resolution Hubble images indicate that condensations A and B (Fig. 5) correspond to places where the spiral arms of NGC 7318b are visibly disrupted: in A, where one of them intersects the shock, one can see a clear detachment of the inner part of this arm dominated by stellar light (emerging from the central bulge) from the gaseous component. This can reasonably be assumed to correspond to a region of higher gas density in the new intruder disk. Higher resolution Chandra data suggest that B is spatially coincident with a relatively compact radio continuum feature (see e.g. Xanthopoulos et al. 2004) near the south end of the shock. An optical spectrum (Gallagher et al. 2001) shows the onset of shock induced line smearing at the same location (with new intruder velocities merging into SQ velocities). A close correspondence between X-ray and H$`\alpha `$+N\[ ii \] emission was already noted in Trinchieri et al. (2003) and is again evident in the comparison with the XMM-Newton data in Fig. 8. Our interference filter image records both H$`\alpha `$ and \[N ii \]$`\lambda `$6583 emission. The association with the shock front is seen only in the H$`\alpha `$+\[N ii \] emission with SQ velocities ($`6600`$ km s<sup>-1</sup>)<sup>1</sup><sup>1</sup>1SQ and new intruder velocity ranges are V$`>`$6000 (mostly 6300-6600) km/s and 5400-6000 km/s respectively (Xu et al. 1999; Sulentic et al. 2001). . Figure 8-left shows two strong concentrations of line emission coincident with the Seyfert nucleus of NGC 7319 and the shock front. There is also evidence of an extended lower surface brightness emission component extending mostly eastward from the shock front that we will discuss later. The multiwavelength morphological similarities are signatures of a close link between different phases of the new intruder ISM and SQ IGM. The 1D intensity cuts along the shock front (Fig. 9, top panels) show more quantitatively the remarkable correspondence between X-ray and line emission. The cuts are centered on A as indicated in Fig. 5. The total extents of the emissions are comparable with a rather sharp boundary at $`50^{\prime \prime }`$ S in all panels. The X rays appear slightly more extended towards the N as more clearly shown with the EPIC-MOS data. A good correspondence between line and X-ray emission is seen between $`30^{\prime \prime }50^{\prime \prime }`$ to the S while the H$`\alpha `$ peak at $`40^{\prime \prime }`$ N, related to starburst A (Xu et al. 1999), does not show a corresponding X-ray feature. The nearby weaker X-ray peak is due to the contribution of component C, which partially enters our cuts but is actually displaced from the position of the H$`\alpha `$ maximum and extends towards the NW. Starburst A is coincident with an X-ray minimum as well as an apparent H i/CO maximum. This probably means that the starburst has occurred recently, with supernova activity still to come. The steep W boundary of the X-ray shock front discovered with Chandra motivated us to suggest a NE to SW transverse component of new intruder motion. Its presence in the X-ray data is difficult to understand in the context of shocks (see Trinchieri et al. 2003). A similarly sharp W boundary of the optical line emission reinforces the significance of this feature and the strong link between X-ray and H$`\alpha `$+\[N ii \] emitting gas. The sharp drop is already evident in the H$`\alpha `$ image (Fig. 8) but is dramatic in the 1D intensity cuts perpendicular to the shock front shown in Fig. 9 (bottom panels). Several examples of sharp X-ray boundaries have now been observed in clusters and groups (i.e. “cold fronts”, Mazzotta et al. 2002; Vikhlinin et al. 2001). However, in these, the spatial discontinuity is thought to correspond to a significant temperature decrease, while SQ does not show this effect. Both drops in X-rays and H$`\alpha `$ could instead correspond to a (pre-existing) density discontinuity. This in turn implies that both emissions derive from impact with the same debris field. It is also conceivable that the X-ray discontinuity corresponds to a contact discontinuity where two separate flows are colliding. In that case, pressure and velocity would be continuous across the surface but the density would suddenly change. Since these discontinuities are generally unstable their presence might be understood if there is a magnetic field present that could suppress Kelvin-Helmholtz type instabilities generated by tangential flows. The detection of a radio continuum emission coincident with the shock front is evidence for the presence of a magnetic field (Williams et al. 2002; Xu et al. 2003). The H$`\alpha `$ boundary at the same position and with a similar drop would need to be explained in the same context. A close connection could also exist between the X-ray and dust distributions. Xu et al. (2003) have emphasized the role of dust in the cooling process behind the shock front resulting from the collision between the intruder galaxy and SQ IGM. They have argued that it could dominate the radiative cooling process. Residual 60$`\mu `$m and 100$`\mu `$m emission detected with ISO (i.e., not attributed to NGC 7319 and the foreground NGC 7320) shows an elongated feature coincident with the ridge observed in H$`\alpha `$, radio continuum and X-ray (see Figure 3 in Xu et al.). Xu et al. (2003) have interpreted this as evidence for shock-related Far InfraRed emission. They derive a sputtering timescale for the dust of a few times $`10^6`$ yr, which is comparable to the timescale for gas cooling due to collision with dust grains. In their model, the expected luminosity for the dust is also comparable to the observed FIR luminosity attributed to the shock region. Since the parameters assumed in Xu et al, based on our Chandra observation, are basically confirmed by the present XMM-Newton data, we do not recalculate these estimates. However, to properly interpret the exact role of the dust emission in the cooling process, and discard other plausible interpretations (e.g. evaluate the possible contribution from the unshocked arm of the new intruder along the same line of sight), a detailed comparison between the dust distribution and the X-ray emission in this region is required, together with a detailed treatment of the shock physics through e.g. numerical simulations. The complexity of the multiphase medium that XMM-Newton data have shown will also have to be taken into account in any detailed modeling. The spectral properties of the shock region are not well defined but, under the assumption of low metal abundance, imply a multi-temperature plasma. This is most likely the observational signature of non-equilibrium effects. Due to the differences in atomic cross-sections for ionization and recombination of different ions and ionization stages of the same ion, their coexistence in a plasma can mimic the signature of a multi-temperature structure. The assumption of equilibrium conditions that is implicit in current plasma models applies when the ionization and recombination time scales of the plasma are short compared to the dynamical and cooling time scales. This is not the case for recently shocked gas. Conversely, if the plasma is expanding rapidly due to its overpressure with respect to the ambient medium, recombination is expected to be delayed (Breitschwerdt & Schmutzler 1994, 1999). Unfortunately, lack of detailed knowledge of the gas dynamics and thermal history in SQ, coupled with the limitations of the data prevents us from applying a proper modeling to account for non-equilibrium conditions. Moreover we cannot discard additional effects due to a inhomogeneous matter distribution and/or geometrical effects. The middle condensation (A) with smallest N<sub>H</sub> could represent the leading edge of the approaching shock, already past the SQ debris field that is visible through H$`\alpha `$+\[N ii \] emission. The other two regions B and C might lag behind and still be embedded in molecular gas responsible for additional absorption. The neutral hydrogen detected in these regions is not enough to explain the inferred absorbing column densities. In fact no H i is detected in the region covered by our ellipse (Williams et al. 2002) although H i features are found at the N and S end with velocities consistent with those in the shock. We have interpreted the H i clouds and NS shock feature as evidence for a preexisting tidal feature possibly produced in earlier tidal activity (Sulentic et al. 2001; Trinchieri et al. 2003). The center of this feature is now shocked as a result of the recent collision with the new intruder. The evidence for a significant amount of debris, traced by the H$`\alpha `$ emission, implies the presence of associated material, that could explain the higher absorption in X-rays. As already discussed in the context of Chandra results (Trinchieri et al. 2003), the temperatures associated with pre- and post-shock medium are a problem in our current understanding of shocks and the physical conditions of SQ: a bow-shock cannot be avoided if the new intruder slams into the debris with supersonic velocity, but the measured temperatures are lower than what is expected from the velocity of the intruder. If the gas ahead of the shock is fairly hot, the shock Mach number is low (M=2-3), resulting in a low compression factor and a moderate increase of the post-shock temperature (2-4 times). With the limitation of the current spectral analysis, we measure a pre-shock temperature of $``$ 0.3-1 keV in “HALO” and “TAIL” that would imply a post-shock temperature in the range 1-4 keV. In shock, we measure kT$`1`$ keV, at the lower end of the expected range (although a hint of higher temperatures are seen in the middle blob, see Table 3). We could of course be observing the system some time after the collision, after significant cooling has already occurred: however the “radiative” cooling times that we derive, $`10^810^9`$ yr, are long compared to the estimated crossing time for NGC7318b and would further imply that the galaxy is now at a distance of several kpc from the site of impact. As suggested in Trinchieri et al. (2003), if the shock is oblique, most of the gas passes through the wings of the bow-shock under some angle and hence compression and heating are reduced. Alternatively, if the gas has been freshly shocked, with ionization lagging behind the entropy increase, the post-shock gas appears to be “under-ionized”, i.e. it has essentially the same spectral characteristics as the unshocked medium. In this non-equilibrium ionization scenario, compression could be high but we could still derive “low temperatures”. In this case, we have to be in a situation where the shock is very recent, because as time goes by the system would gradually be getting back to equilibrium conditions. If cooling by dust is a contributing or even dominant process, then the timescale for the shock could be extremely short, down to a few $`\times 10^6`$ yr (Xu et al. 2003). In this case, if the cooling and sputtering times are comparable, the process could be self-regulating and result in an almost iso-thermal shock, explaining the lack of a temperature increase in the region. ### 3.3 Separating Shocked and Diffuse X-ray Components in SQ The analogies that compact groups show to rich clusters (e.g. extreme densities with resultant strong galaxy harassment) have led to claims that compact groups also typically show diffuse X-ray haloes. If primordial or at least virialized they offer an important chance to map the gravitational potential of the aggregate. Several examples of X-ray clouds are found in compact groups: Hickson 62 represents probably the most unambiguous example of such a component but lower luminosity and less massive examples have been found in other compact groups mostly with low spiral fraction (Mulchaey et al. 2003; Osmond & Ponman 2004). In the case of SQ, our inferences about any long-lived diffuse component have been clouded by the strong and complex emission from the large scale shock. Low resolution ROSAT-PSPC and ASCA observations were interpreted as emission dominated by a diffuse component (Sulentic et al. 1995; Awaki et al. 1997). Subsequent ROSAT-HRI and Chandra observations (Pietsch et al. 1997; Sulentic et al. 2001; Trinchieri et al. 2003) showed that most of the diffuse photons originated from the extended shock leaving little flux that could be ascribed to diffuse emission. XMM-Newton observations reported here provide the strongest evidence for diffuse emission in SQ because the observed extent of the X-ray emission exceeds any reasonable estimate for the extent of shock related activity. Spectral data for the extended emission also require a multi-component plasma model in analogy with the shock region. In fact there seems to be a continuity in the spectral properties out to the TAIL region. The surrounding region is apparently inhomogeneous like the shock front with a highly variable N<sub>H</sub> column and an unrelaxed diffuse component mixed with emission related to the ongoing shock. The complex H i distribution in SQ may create or enhance the observed inhomogeneities. There are two X-ray components that we would like to separate: 1) shock related emission, likely concentrated near the NS feature, where however emission related to the shock proper, the new intruder disk and a more long-lived diffuse component are probably mixed, and 2) the more long-lived diffuse component itself, of lower surface brightness and larger distribution. We will make use of the narrow band H$`\alpha `$+\[N ii \] and broad R band images to respectively infer the extent of shock and diffuse emission components. The 1D EW cuts shown in Fig. 9 indicate a significant extension towards the E in both line and X-ray emissions. The eastward extension is more significant in the new XMM-Newton data and can be traced up to $`6070^{\prime \prime }`$ ($``$40kpc) east from the shock front and enveloping the Seyfert nucleus of NGC 7319. A similar diffuse extension up to $``$ 50<sup>′′</sup> is observed in line emission. Emission westward of the NS shock front is also visible in X-ray (mostly XMM-Newton) data, but at a lower intensity (see also Fig. 5 in Trinchieri et al. (2003)). H$`\alpha `$+\[N ii \] cannot be reliably mapped here because of problems associated with subtraction of the continuum emission near the centers of NGC 7318ab. The simplest interpretation sees the eastern emission extension as related to the new intruder disk. It may be the signature of shocked eastern half of the disk which is reasonable because almost no signature of neutral or warm gas is found eastward of the shock front indicating that the disk has already passed the front. On the contrary, two H i and numerous H ii regions with new intruder velocities are observed west of the shock front. This east-west difference suggests that either the western side of the new intruder disk encountered much less SQ gas in its path or that it has not yet passed through SQ. The former appears likely because the H i and H ii associated with the western side indicate that the disk is disrupted. The high line of sight velocity of the intruder makes this disruption difficult to explain with a response time of t$`<`$10<sup>8</sup> years. It is difficult to infer a total luminosity for the shock related component. The bulk of the shocked emission is inferred to follow the contours superimposed on Figure 8b where diffuse (assumed to be forbidden \[N ii \]) line emission is found. It is the most reasonable approach we can take given the complexity of the source and it is reasonable to say that it is on the order of $`10^{41}`$ erg s<sup>-1</sup> which would be high for a normal spiral unless enhanced by a collision. Low level emission west of the shock front as well as extended emission towards the SE (TAIL) and SW (towards NGC 7317) are the strongest evidences for large scale diffuse X-ray emission in SQ (see Fig.1 and Fig.3). The irregular shape of this component suggests that it is far from dynamically relaxed. An optical or an X-ray diffuse halo will grow by sequential stripping in a compact group like SQ. Earlier analysis of optical data (Moles et al. 1998) showed a halo luminosity L$``$L and suggested a smoother distribution of diffuse light towards the west and a very complex distribution towards the east. The latter was interpreted as evidence for the most recent halo building event connected with the last passages of NGC 7320c and stripping of NGC 7319. The halo building component involves the optical tidal tails and associated debris between NGC 7319 and NGC 7320c. Star formation condensations have also been found in this debris many tens of kpc from group members/intruders (Sulentic et al. 2001; Mendes et al. 2005). Recent ultra-deep CCD imagery (Gutiérrez et al. 2002) considerably expands the size of the detected optical halo reinforcing the earlier conclusion that it represents an even more significant luminosity, and hence baryonic mass, component than previously estimated. The 2D extent of the diffuse emission in both X-ray and optical light can be inferred from Fig. 1 and the comparison shown in Fig. 4 (X-ray contours on red image, left, and optical contours on X-ray color image, right). There is again a striking correspondence between diffuse emissions from hot gas and the stellar envelope in the southern half. Incidentally, this makes a strong case for the dynamical involvement of NGC 7317 in SQ (see Moles et al. 1998). An extension towards NNE is also seen in both X-ray and optical light (however, beware of the prominent Seyfert nucleus in the X-ray image). This extended emission cannot reasonably be ascribed to the ongoing shock and was apparently produced in earlier collisions within SQ that are hard to reconstruct now. The lobe-like structures are likely signatures of particular stripping episodes that have not had time to relax. The latest stripping episode involving debris towards the old intruder is even less relaxed and therefore more directly traceable to the last one or two intrusions of NGC 7320c. SQ is apparently an example of extensive diffuse light in a dynamically active compact group. Fig. 10 shows a further comparison between X-ray and red continuum emission intended to show the extent of diffuse emission towards the SW and NE directions. The 1D track is also shown in Fig 5 which is centered roughly SE of feature C. Unlike Fig. 9, the shapes of the two distributions are not similar. The “two-horned” profile in red continuum light is due to NGC 7318 while the peak in the X-ray plot is due to feature C and the bridge that connects it to the NS shock front. However diffuse emission can be traced in both plots out to at least to 200<sup>′′</sup> ( 70kpc) SW and $`100^{\prime \prime }`$ to the NE. This again indicates that the extended optical halo shows a correspondence with the extended X-ray halo. It is difficult to say more about the multiphase IGM of SQ especially in the absence of 3D data. An estimate for the X-ray luminosity of the (non-shock) diffuse emission yields L$`{}_{x}{}^{}5\times 10^{41}`$ erg s<sup>-1</sup> (0.5-2 keV) comparable with the shocked component but of much lower surface brightness. It is also comparable to luminosities of the hot gaseous component in other low velocity dispersion groups discussed by Helsdon et al. (2005). SQ is a very low velocity dispersion group if the new intruder not included in the estimate. ## 4 Conclusions Chandra and XMM-Newton have provided complimentary insights for the complex X-ray emission from Stephan’s Quintet. The strongest point source involves the Seyfert nucleus of NGC 7319 although all members have now been detected, except for the two intruders NGC 7320c and NGC 7318b. The most unusual aspect of the X-ray emission from SQ involves its extended emission. All of the galaxies in SQ have been stripped by a continuing sequence of intrusions by neighboring and likely member galaxies. The most recently stripped spiral galaxies involve NGC 7319 and NGC 7320c. Apparently enough gas was channeled into the nucleus of the former to stimulate an AGN while the latter is undetected at all but optical wavelengths. The stripping events have given rise to a complex multiphase medium that has now been detected in radio continuum, radio lines (H i+CO), optical line (H$`\alpha `$+\[N ii \]) and X-ray. Much/most of the radio continuum, optical line and X-ray emission are connected with a large scale shock generated by the most recent collision involving the ISM of NGC 7318b and the preexisting IGM debris field from previous stripping events. We infer an extent for the shock dominated emission in excess of D$``$50kpc. The stronger emission involves a narrow NS shock front that may reflect the existence of an extended debris field connecting two H i clouds in the new intruder path. The more extended shock related emission is interpreted as the shocked new intruder disk. The 2D correspondence between the X-ray and optical line emission suggests a common origin for these two gas phases. We now detect a much more extended (D 130-150kpc) diffuse component that cannot be ascribed to the shock. This is presumably related to the diffuse components found in other groups and clusters. No line emission maps this X-ray structure, but rather diffuse light from a halo of stripped stars. The complex lobe-like structure of this halo presumably reflects a group far from a state of dynamical relaxation. It is difficult to find an analogy to SQ in the current X-ray literature. The closest might be the Antennae (Fabbiano et al. 2003; Metz et al. 2004) where strong interactions have obviously taken place. However the analogy quickly breaks down, because the Antennae are dominated by star formation regions and most of the diffuse X-ray emission is associated with these regions. In addition, while strongly distorted, most of the gas still resides in the galaxies. We do not see evidence for widespread star formation activity in SQ galaxies, although we see small star forming condensations developing in the stripped debris (Xu et al. 1999; Sulentic et al. 2001; Mendes et al. 2005). Shocked emission in SQ does not follow these few features: it is concentrated in a more localized region presumably at the sight of current impact and appears to be more extended than traced by the H$`\alpha `$ emission (although mixing with the pre-existing diffuse emission might play a role). The spectra of the diffuse components in SQ and the Antennae both require multi-temperature fits, however it is very likely that this is just the observational signature of very complex temperature distributions and/or non-equilibrium effects. In the Antennae, a detailed analysis has shown the presence of regions at different temperatures and absorption depths (Fabbiano et al. 2003). In SQ we both have evidence of condensations at different temperatures and expect non-equilibrium conditions. Unfortunately, current observations do not allow us to calculate the dynamics and the thermal history of the plasma self-consistently, to properly model its physical conditions. ###### Acknowledgements. We thank the referee, Dr. C. Xu, for pointing out to us the relevance of dust cooling in the shock region, and for several useful comments that helped improve the paper. This research has made use of SAOImage DS9, developed by the Smithsonian Astrophysical Observatory. GT acknowledges partial financial support by the Italian Space Agency ASI. JS acknowledge NASA support under contract NAG5-11204. We thank Carlo Gutierrez for making his optical R-band data available.
warning/0506/math0506493.html
ar5iv
text
# Characterizations of compact and discrete quantum groups through second duals ## Introduction Recently, J. Kustermans and S. Vaes introduced a surprisingly simple set of axioms for what they call *locally compact quantum groups* (\[K–V 1\] and \[K–V 2\]). These axioms cover both the Kac algebras (\[E–S\])—and thus all locally compact groups—as well as the compact quantum groups in the sense of \[Wor\] and allow for a notion of duality that extends the Pontryagin duality for locally compact abelian groups. What makes the locally compact quantum groups interesting from the point of view of abstract harmonic analysis is that many results in abstract harmonic analysis can easily be reformulated in the language of locally compact quantum groups. For instance, Leptin’s theorem (\[Lep\]) asserts that a locally compact compact group $`G`$ is amenable if and only if the Fourier algebra $`A(G)`$ (\[Eym\]) has a bounded approximate identity. This statement can be rephrased as: $`G`$ is amenable if and only if its quantum group dual is co-amenable (as defined in \[B–T, Definition 3.1\]). Indeed, the “if” part is true for any locally compact quantum group (\[B–T, Theorem 3.2\]) whereas the converse is known to be true only in the discrete case (\[Tom\]; see also \[Rua\] and \[B–T\]). In \[Wat\] (see also \[Gro\] and \[Joh\]), S. Watanabe proved that a locally compact group $`G`$ is compact if and only if its group algebra $`L^1(G)`$ is an ideal in its second dual. On the other hand, A. T.-M. Lau showed that $`A(G)`$ is an ideal in $`A(G)^{}`$ if and only if $`G`$ is discrete (\[Lau, Theorem 3.7\]). We shall see that both results are just two facets of one theorem on locally compact quantum groups. To this end, note that, according to \[K–V 2\], one way of describing a locally compact quantum group is as a Hopf–von Neumann algebra $`(𝔐,\mathrm{\Gamma })`$ with additional structure. Hence, the predual $`𝔐_{}`$ of $`𝔐`$ is a Banach algebra in a canonical way. Simultaneously extending the results by both Watanabe and Lau, we shall prove that the locally compact quantum group $`(𝔐,\mathrm{\Gamma })`$ is compact if and only if $`𝔐_{}`$ is an ideal in its second dual. If $`G`$ is a discrete group, then $`𝒞_0(G)=c_0(G)`$ is an ideal in $`𝒞_0(G)^{}=\mathrm{}^{\mathrm{}}(G)`$. On the other hand, if $`𝒞_0(G)`$ is an ideal in $`𝒞_0(G)^{}`$, then multiplication in $`𝒞_0(G)`$ is weakly compact (\[Pal, Proposition 1.4.13\]) and thus compact (because $`𝒞_0(G)`$ has the Dunford–Pettis property), so that $`G`$ is discrete. We shall see that this result also extends to locally compact quantum groups: if $`(𝔄,\mathrm{\Gamma })`$ is a reduced $`C^{}`$-algebraic quantum group, then $`(𝔄,\mathrm{\Gamma })`$ is discrete if and only if $`𝔄`$ is an ideal in $`𝔄^{}`$. ## 1 Compactness and amenability for Hopf–von Neumann algebras In this section, we formulate and briefly discuss the notions of compactness and amenability in a general Hopf–von Neumann algebra context. We begin with recalling the definition of a Hopf–von Neumann algebra ($`\overline{}`$ denotes the $`W^{}`$-tensor product): ###### Definition 1.1. A *Hopf–von Neumann algebra* is a pair $`(𝔐,\mathrm{\Gamma })`$, where $`𝔐`$ is a von Neumann algebra and $`\mathrm{\Gamma }:𝔐𝔐\overline{}𝔐`$ is a *co-multiplication*, i.e., a normal, unital -homomorphism satisfying $`(\mathrm{id}\mathrm{\Gamma })\mathrm{\Gamma }=(\mathrm{\Gamma }\mathrm{id})\mathrm{\Gamma }`$. The main examples we have in mind are from abstract harmonic analysis: ###### Examples. 1. For a locally compact group $`G`$, define $`\mathrm{\Gamma }_G:L^{\mathrm{}}(G)L^{\mathrm{}}(G\times G)`$ by letting $$(\mathrm{\Gamma }_G\varphi )(x,y):=\varphi (xy)(\varphi L^{\mathrm{}}(G),x,yG).$$ Then $`(L^{\mathrm{}}(G),\mathrm{\Gamma }_G)`$ is a Hopf–von Neumann algebra. 2. For a locally compact group $`G`$, let $`\lambda `$ denote its left regular representation, and let $`\mathrm{VN}(G):=\lambda (G)^{\prime \prime }`$ denote the *group von Neumann algebra* of $`G`$. Then $$\widehat{\mathrm{\Gamma }}_G:\mathrm{VN}(G)\mathrm{VN}(G)\overline{}\mathrm{VN}(G),\lambda (x)\lambda (x)\lambda (x)$$ defines a co-multiplication, so that $`(\mathrm{VN}(G),\widehat{\mathrm{\Gamma }}_G)`$ is a Hopf–von Neumann algebra. Given a Hopf–von Neumann algebra $`(𝔐,\mathrm{\Gamma })`$, one can define a product $``$ on $`𝔐_{}`$, the unique predual of $`𝔐`$, turning it into a Banach algebra: $$(fg)(x):=(fg)(\mathrm{\Gamma }x)(f,g𝔐_{},x𝔐).$$ (1) ###### Example. Let $`G`$ be a locally compact group. Then applying (1) to $`(L^{\mathrm{}}(G),\mathrm{\Gamma }_G)`$ yields the usual convolution product on $`L^1(G)`$ whereas, for $`(\mathrm{VN}(G),\widehat{\mathrm{\Gamma }}_G)`$, we obtain pointwise multiplication on $`A(G)`$ (\[Eym\]). For any Banach algebra $`𝔄`$, there are two canonical ways to extend the product to the second dual: the two *Arens products* (see \[Dal\] or \[Pal\]). These two products need not coincide; they are identical, however, whenever one of the two factors involved is from $`𝔄`$. Given a Hopf–von Neumann algebra $`(𝔐,\mathrm{\Gamma })`$, we will write $``$ for any of the two Arens products on $`𝔐_{}^{}=𝔐^{}`$. We now define what it means for a Hopf–von Neumann algebra to be (left) amenable and compact, respectively: ###### Definition 1.2. Let $`(𝔐,\mathrm{\Gamma })`$ be a Hopf–von Neumann algebra. Then we call $`(𝔐,\mathrm{\Gamma })`$ *left amenable* if there is a *left invariant* state $`M𝔐^{}`$, i.e., satisfying $$fM=f(1)M(f𝔐_{}).$$ If $`M`$ can be chosen to be in $`𝔐_{}`$, we call $`(𝔐,\mathrm{\Gamma })`$ *left compact*. ###### Example. If $`G`$ is a locally compact group, then $`(L^{\mathrm{}}(G),\mathrm{\Gamma }_G)`$ is left amenable if and only if $`G`$ is amenable and left compact if and only if $`G`$ is compact. Of course one can equally well define right amenable (compact) Hopf–von Neumann algebras—through the existence of right invariant (normal) states—as well as amenable (compact) Hopf–von Neumann algebras by demanding the existence of a (normal) state that is both left and right invariant. The following is well known, but for convenience, we include a proof: ###### Proposition 1.3. Let $`(𝔐,\mathrm{\Gamma })`$ be a Hopf–von Neumann algebra. Then $`(𝔐,\mathrm{\Gamma })`$ is amenable (compact) if and only if $`(𝔐,\mathrm{\Gamma })`$ is both left and right amenable (compact). ###### Proof. If $`M_lM^{}`$ is a left invariant state and $`M_r𝔐^{}`$ is a right invariant state, then $`M_lM_r`$ is an invariant state (no matter which Arens product is chosen on $`𝔐^{}`$). Moreover, if $`M_l`$ and $`M_r`$ are both normal, then so is $`M_lM_r`$. ∎ We conclude this section with recalling the notion of a co-involution. ###### Definition 1.4. Let $`(𝔐,\mathrm{\Gamma })`$ be a Hopf–von Neumann algebra. A *co-involution* for $`(𝔐,\mathrm{\Gamma })`$ is a -antihomomorphism $`R:𝔐𝔐`$ with $`R^2=\mathrm{id}`$ satisfying $$(RR)\mathrm{\Gamma }=\sigma \mathrm{\Gamma }R,$$ where $`\sigma `$ is the flip map on $`𝔐\overline{}𝔐`$. A co-involution $`R`$ for a Hopf–von Neumann algebra $`(𝔐,\mathrm{\Gamma })`$ is necessarily normal and can be used to define an involution on the Banach algebra $`𝔐_{}`$. For $`f𝔐_{}`$, define $`\overline{f}𝔐_{}`$ by letting $$\overline{f}(x):=\overline{f(x^{})}(x𝔐),$$ and set $`f^{\mathrm{}}:=\overline{f}R`$. It is easily seen that is indeed an involution. In view of Proposition 1.3, we obtain: ###### Corollary 1.5. The following are equivalent for a Hopf–von Neumann algebra $`(𝔐,\mathrm{\Gamma })`$ with co-involution: 1. $`(𝔐,\mathrm{\Gamma })`$ is left amenable (compact); 2. $`(𝔐,\mathrm{\Gamma })`$ is right amenable (compact); 3. $`(𝔐,\mathrm{\Gamma })`$ is amenable (compact). ## 2 Locally compact quantum groups Compact quantum groups in the $`C^{}`$-algebraic setting were defined by S. L. Woronowicz (see \[Wor\] and \[M–vD\] for accounts). In \[K–V 1\], J. Kustermans and S. Vaes introduced a comparatively simple set of axioms to define general locally compact quantum groups in a $`C^{}`$-algebraic context. Alternatively, locally compact quantum groups can also be described as Hopf–von Neumann algebras with additional structure (see \[K–V 2\] and \[vDae\]): both approaches are equivalent. In this section, we give an outline of the von Neumann algebraic approach to locally compact quantum groups. For details, see \[K–V 1\], \[K–V 2\], and also \[Kus\]. We start with recalling some notions about weights on von Neumann algebras (see \[Tak 2\], for instance). Let $`𝔐`$ be a von Neumann algebra, and let $`𝔐^+`$ denote its positive elements. A *weight* on $`𝔐`$ is an additive map $`\varphi :𝔐^+[0,\mathrm{}]`$ such that $`\varphi (tx)=t\varphi (x)`$ for $`t[0,\mathrm{})`$ and $`x𝔐^+`$. We let $$_\varphi ^+:=\{x𝔐^+:\varphi (x)<\mathrm{}\},_\varphi :=\text{the linear span of }_\varphi ^+,$$ and $$𝒩_\varphi :=\{x𝔐:x^{}x_\varphi \}.$$ Then $`\varphi `$ extends to a linear map on $`_\varphi `$, and $`𝒩_\varphi `$ is a left ideal of $`𝔐`$. Using the GNS-construction (\[Tak 2, p. 42\]), we obtain a representation $`\pi _\varphi `$ of $`𝔐`$ on some Hilbert space $`_\varphi `$; we denote the canonical map from $`𝒩_\varphi `$ into $`_\varphi `$ by $`\mathrm{\Lambda }_\varphi `$. Moreover, we call $`\varphi `$ *finite* if $`_\varphi ^+=𝔐^+`$, *semi-finite* if $`_\varphi `$ is $`w^{}`$-dense in $`𝔐`$, *faithful* if $`\varphi (x)=0`$ for $`x𝔐^+`$ implies that $`x=0`$, and *normal* if $`sup_\alpha \varphi (x_\alpha )=\varphi \left(sup_\alpha x_\alpha \right)`$ for each bounded, increasing net $`(x_\alpha )_\alpha `$ in $`𝔐^+`$. If $`\varphi `$ is faithful and normal, then the corresponding representation $`\pi _\varphi `$ is faithful and normal, too (\[Tak 2, Proposition VII.1.4\]). ###### Definition 2.1. A *locally compact quantum group* is a Hopf–von Neumann algebra $`(𝔐,\mathrm{\Gamma })`$ such that: 1. there is a normal, semifinite, faithful weight $`\varphi `$ on $`𝔐`$—a *left Haar weight*—which is left invariant, i.e., satisfies $$\varphi ((f\mathrm{id})(\mathrm{\Gamma }x))=f(1)\varphi (x)(f𝔐_{},x_\varphi );$$ 2. there is a normal, semifinite, faithful weight $`\psi `$ on $`𝔐`$—a *right Haar weight*—which is right invariant, i.e., satisfies $$\varphi ((\mathrm{id}f)(\mathrm{\Gamma }x))=f(1)\psi (x)(f𝔐_{},x_\psi ).$$ ###### Example. Let $`G`$ be a locally compact group. Then the Hopf–von Neumann algebra $`(L^{\mathrm{}}(G),\mathrm{\Gamma }_G)`$ is a locally compact quantum group: $`\varphi `$ and $`\psi `$ can be chosen as left and right Haar measure, respectively. Even though only the existence of a left and a right Haar weight, respectively, is presumed, both weights are actually unique up to a positive scalar multiple (see \[K–V 1\] and \[K–V 2\]). For each locally compact quantum group $`(𝔐,\mathrm{\Gamma })`$, there is a unique unitary—the *multiplicative unitary*$`W(_\varphi _2_\varphi )`$, where $`_2`$ stands for the Hilbert space tensor product, such that $$W^{}(\mathrm{\Lambda }_\varphi (x)\mathrm{\Lambda }_\varphi (y))=(\mathrm{\Lambda }_\varphi \mathrm{\Lambda }_\varphi )((\mathrm{\Gamma }y)(x1))(x,y𝒩_\varphi )$$ (\[K–V 2, Theorem 1.2\]). The unitary $`W`$ lies in $`𝔐\overline{}(_\varphi )`$ and implements the co-multiplication via $$\mathrm{\Gamma }x=W^{}(1x)W(x𝔐)$$ (see the discussion following \[K–V 2, Theorem 1.2\]). As discussed in \[K–V 1\] and \[K–V 2\], locally compact quantum groups can equivalently be described in $`C^{}`$-algebraic terms. The corresponding $`C^{}`$-algebra is obtained as $$𝔄:=\overline{\{(\mathrm{id}\nu )(W):\nu (_\varphi )_{}\}}^{}.$$ To emphasize the parallels between locally compact quantum groups and groups, we shall use the following notation (which was suggested by Z.-J. Ruan): a locally compact quantum group $`(𝔐,\mathrm{\Gamma })`$ is denoted by the symbol $`𝔾`$, and we write $`L^{\mathrm{}}(𝔾)`$ for $`𝔐`$, $`L^1(𝔾)`$ for $`𝔐_{}`$, $`L^2(𝔾)`$ for $`_\varphi `$, and $`𝒞_0(𝔾)`$ for $`𝔄`$. If $`L^{\mathrm{}}(𝔾)=L^{\mathrm{}}(G)`$ for a locally compact group $`G`$ and $`\mathrm{\Gamma }=\mathrm{\Gamma }_G`$, we say that $`𝔾`$ actually *is* a locally compact group. The *left regular* representation of a locally compact quantum group $`𝔾`$ is defined as $$\lambda :L^1(𝔾)(L^2(𝔾)),f(f\mathrm{id})(W).$$ (If $`𝔾`$ is a locally compact group $`G`$, this is just the usual left regular representation of $`L^1(G)`$ on $`L^2(G)`$.) It is a faithful, contractive representation of the Banach algebra $`L^1(𝔾)`$ on $`L^2(𝔾)`$. Locally compact quantum groups allow for the development of a duality theory that extends Pontryagin duality for locally compact abelian groups. Set $$L^{\mathrm{}}(\widehat{𝔾}):=\overline{\lambda (L^1(𝔾))}^{\sigma \text{-strongly}\text{}}.$$ Then $`L^{\mathrm{}}(\widehat{𝔾})`$ is a von Neumann algebra, and $$\widehat{\mathrm{\Gamma }}:L^{\mathrm{}}(\widehat{𝔾})L^{\mathrm{}}(\widehat{𝔾})\overline{}L^{\mathrm{}}(\widehat{𝔾}),x\sigma W(x1)W^{}\sigma $$ is a co-multiplication (here, $`\sigma `$ is the flip map on $`L^2(𝔾)_2L^2(𝔾)`$). One can also define a left Haar weight $`\widehat{\varphi }`$ and a right Haar weight $`\widehat{\psi }`$ for $`(L^{\mathrm{}}(\widehat{𝔾}),\widehat{\mathrm{\Gamma }})`$ turning it into a locally compact quantum group again, the *dual quantum group* of $`𝔾`$, which we denote by $`\widehat{𝔾}`$. Generally, if $`X`$ is an object associated with $`𝔾`$, we convene to denote the corresponding object associated with $`\widehat{𝔾}`$ by $`\widehat{X}`$. Finally, a Pontryagin duality theorem holds, i.e., $`\widehat{\widehat{𝔾}}=𝔾`$. ###### Example. If $`𝔾`$ is a locally compact group $`G`$, then $`\widehat{𝔾}=(\mathrm{VN}(G),\widehat{\mathrm{\Gamma }}_G)`$, and $`\widehat{\varphi }=\widehat{\psi }`$ is the Plancherel weight on $`\mathrm{VN}(G)`$ (\[Tak 2, Definition VII.3.2\]). Each locally compact quantum group $`𝔾`$ is equipped with a canonical co-involution—the *unitary antipode*—, so that $`L^1(𝔾)`$ can be equipped with an involution . Unlike for groups, however,—and more generally for Kac algebras (see \[E–S\])—the left regular representation of $`L^1(𝔾)`$ need not be a -representation with respect to this involution. The *antipode* of $`𝔾`$ is a $`\sigma `$-strongly-closed operator $`S`$ on $`L^{\mathrm{}}(𝔾)`$ whose domain $`𝒟(S)`$ is $`\sigma `$-strongly-dense. In general, $`S`$ is not totally defined. Letting $$L_{}^1(𝔾):=\{fL^1(𝔾):\text{there is }gL^1(𝔾)\text{ such that }g(x)=\overline{f}(Sx)\text{ for }x𝒟(x)\},$$ we obtain a dense subalgebra of $`L^1(𝔾)`$. On $`L_{}^1(𝔾)`$, we can then define an involution by letting $$f^{}(x):=\overline{f}(Sx)(x𝒟(S)).$$ Defining a norm $`||||||`$ on $`L_{}^1(𝔾)`$ via $$|f|:=\mathrm{max}\{f,f^{}\}(fL_{}^1(𝔾)),$$ we turn $`L_{}^1(𝔾)`$ into a Banach -algebra (in fact, with isometric involution). The restriction of $`\lambda `$ to $`L_{}^1(𝔾)`$ is then a -representation of $`L_{}^1(𝔾)`$. Furthermore, as a Banach -algebra, $`L_{}^1(𝔾)`$ has an enveloping $`C^{}`$-algebra, which we denote by $`𝒞_0^u(\widehat{𝔾})`$. For more details, see \[Kus\] ## 3 Compact quantum groups We shall call a locally compact quantum group $`𝔾`$ *compact* if the Hopf–von Neumann algebra $`(L^{\mathrm{}}(𝔾),\mathrm{\Gamma })`$ is (left) compact in the sense of Definition 1.2. Other characterizations of compactness for locally compact quantum groups are collected below: ###### Proposition 3.1. The following are equivalent for a locally compact quantum group $`𝔾`$: 1. $`𝔾`$ is compact; 2. the $`C^{}`$-algebra $`𝒞_0(𝔾)`$ is unital; 3. the left Haar weight of $`𝔾`$ is finite. ###### Proof. (i) $``$ (ii) is \[B–T, Proposition 3.1\] (compare also \[Tom, Remark 3.8(5)\]). (ii) $``$ (iii): Let $`\varphi `$ denote the left Haar weight of $`𝔾`$. Since $`𝒞_0(𝔾)𝒩_\varphi `$ is a norm dense left ideal of $`𝒞_0(𝔾)`$ is must contain the identity, so that $`\varphi `$ is finite. (iii) $``$ (i) is trivial. ∎ In this section, we shall prove another, less straightforward characterization of the compact quantum groups: a locally compact quantum group $`𝔾`$ is compact if and only if $`L^1(𝔾)`$ is an ideal in $`L^1(𝔾)^{}`$. We first prove a lemma for general locally compact quantum groups: ###### Lemma 3.2. Let $`𝔾`$ be a locally compact quantum group, and let $`fL_{}^1(𝔾)`$. Then we have $$\lambda (f^{})\mathrm{\Lambda }_\varphi (x)=\mathrm{\Lambda }_\varphi ((\overline{f}\mathrm{id})(\mathrm{\Gamma }x))(x𝒩_\varphi ).$$ ###### Proof. Fix $`x𝒩_\varphi `$, and let $`\nu (L^2(𝔾))_{}`$ be arbitrary. Define $`x\nu (L^2(𝔾))_{}`$ by letting $`(x\nu )(T)=\nu (Tx)`$ for $`T(L^2(𝔾))`$. We obtain that $$\begin{array}{cc}\hfill \nu (\lambda (f^{})x)& =(f^{}\nu )(W(1x))\hfill \\ & =(f^{}x\nu )(W)\hfill \\ & =\overline{f}(S(\mathrm{id}x\nu )(W)))\hfill \\ & =\overline{f}((\mathrm{id}x\nu )(W^{})),\text{by }\text{[K–V 1, Proposition 8.3]},\hfill \\ & =(\overline{f}\nu )(W^{}(1x))\hfill \\ & =\nu ((\overline{f}\mathrm{id})(W^{})x).\hfill \end{array}$$ Since $`\nu (L^2(𝔾))_{}`$ is arbitrary, this means that $$\lambda (f^{})x=(\overline{f}\mathrm{id})(W^{})x.$$ In view of the fact that $`x𝒩_\varphi `$, it follows from \[K–V 1, (8.3)\] that $$\lambda (f^{})\mathrm{\Lambda }_\varphi (x)=(\overline{f}\mathrm{id})(W^{})\mathrm{\Lambda }_\varphi (x)=\mathrm{\Lambda }_\varphi ((\overline{f}\mathrm{id})(\mathrm{\Gamma }x)),$$ as claimed. ∎ Suppose now that $`𝔾`$ is a compact quantum group. Since the left Haar weight $`\varphi `$ is then finite by Proposition 3.1—and can be chosen to be a state—, $`\mathrm{\Lambda }_\varphi :𝒩_\varphi L^2(𝔾)`$ is a contraction defined on all of $`L^{\mathrm{}}(𝔾)`$. Define $`\iota :L^2(𝔾)L^1(𝔾)`$ by letting $$\iota (\xi )(x)=\xi ,\mathrm{\Lambda }_\varphi (x^{})(\xi L^2(𝔾),xL^{\mathrm{}}(𝔾)).$$ Clearly, $`\iota `$ is a linear, injective contraction with dense range. The following compatibility result holds: ###### Lemma 3.3. Let $`𝔾`$ be a compact quantum group. Then we have $$f\iota (\xi )=\iota (\lambda (f)\xi )(fL^1(𝔾),\xi L^2(𝔾)).$$ (2) ###### Proof. By Lemma 3.2 and the fact that $`\lambda `$ is a -representation of $`L_{}^1(𝔾)`$, we have $$\lambda (f)^{}\mathrm{\Lambda }_\varphi (x)=\mathrm{\Lambda }_\varphi ((\overline{f}\mathrm{id})(\mathrm{\Gamma }x))(fL_{}^1(𝔾),xL^{\mathrm{}}(𝔾)).$$ (3) Hence, we thus obtain $$\begin{array}{cc}\hfill \iota (\lambda (f)\xi )(x)& =\lambda (f)\xi ,\mathrm{\Lambda }_\varphi (x^{})\hfill \\ & =\xi ,\lambda (f)^{}\mathrm{\Lambda }_\varphi (x^{})\hfill \\ & =\xi ,\mathrm{\Lambda }_\varphi ((\overline{f}\mathrm{id})(\mathrm{\Gamma }x^{})),\text{by (}\text{3}\text{)},\hfill \\ & =\iota (\xi )(((\overline{f}\mathrm{id})(\mathrm{\Gamma }x^{}))^{})\hfill \\ & =\iota (\xi )((f\mathrm{id})(\mathrm{\Gamma }x))\hfill \\ & =(f\iota (\xi ))(\mathrm{\Gamma }x)\hfill \\ & =(f\iota (\xi ))(x)(fL_{}^1(𝔾),\xi L^2(𝔾),xL^{\mathrm{}}(𝔾))\hfill \end{array}$$ and thus $$\iota (\lambda (f)\xi )=f\iota (\xi )(fL_{}^1(𝔾),\xi L^2(𝔾)).$$ Since $`L_{}^1(𝔾)`$ is dense in $`L^1(𝔾)`$, this yields (2). ∎ Let $`E`$ and $`F`$ be Banach spaces, and let $`T:EF`$ be linear. Recall that $`T`$ is said to be *weakly compact* if it maps the unital ball of $`E`$ onto a relatively weakly compact subset of $`F`$. As a consequence of Lemma 3.3, we obtain: ###### Proposition 3.4. Let $`𝔾`$ be a compact quantum group. Then multiplication in $`L^1(𝔾)`$ is weakly compact. ###### Proof. Let $$I:=\{gL^1(𝔾):L^1(𝔾)ffg\text{ is weakly compact}\}.$$ We claim that $`I=L^1(𝔾)`$. To see this, note first that—due to the reflexivity of $`L^2(𝔾)`$—the map $$L^1(𝔾)L^2(𝔾),f\lambda (f)\xi $$ is trivially weakly compact for each $`\xi L^2(𝔾)`$. Since $`\iota :L^2(𝔾)L^1(𝔾)`$ is continuous, Lemma 3.3 yields that $`\iota (L^2(𝔾))I`$. Since $`I`$ is closed and $`\iota (L^2(𝔾))`$ is dense in $`L^1(𝔾)`$, we conclude that $`I=L^1(𝔾)`$. Since $$fg=(g^{\mathrm{}}f^{\mathrm{}})^{\mathrm{}}(f,gL^1(𝔾)),$$ it follows that multiplication—both from the left and from the right—is indeed weakly compact in $`L^1(𝔾)`$. ∎ It is a well known Banach algebraic fact that a Banach algebra is an ideal in its second dual if and only if multiplication in $`𝔄`$ is weakly compact (\[Pal, Proposition 1.4.13\]). Hence, we obtain: ###### Corollary 3.5. Let $`𝔾`$ be a compact quantum group. Then $`L^1(𝔾)`$ is an ideal in $`L^1(𝔾)^{}`$. To prove the converse of Corollary 3.5, we first prove two lemmas, which are patterned on \[M–vD, Lemmas 4.2 and 4.3\]. (We are grateful to S. Vaes for outlining this argument.) ###### Lemma 3.6. Let $`𝔾`$ be a locally compact quantum group such that $`L^1(𝔾)`$ is an ideal in $`L^1(𝔾)^{}`$. Then, for each state $`fL^1(𝔾)`$, there is a state $`gL^1(𝔾)`$ such that $`fg=gf=g`$. ###### Proof. Let $`gL^1(𝔾)^{}`$ be a $`\sigma (L^1(𝔾)^{},L^{\mathrm{}}(𝔾))`$-accumulation point of the sequence $`\left(\frac{1}{n}_{k=1}^nf^k\right)_{n=1}^{\mathrm{}}`$, where $`f^k`$ stands for the $`k`$-th power of $`f`$ with respect to the product $``$ in $`L^1(𝔾)`$. An inspection of the proof of \[M–vD, Lemma 4.2\] yields that $`gL^1(𝔾)^{}`$ is a state satisfying $`fg=gf=g`$. Since $`L^1(𝔾)`$ is an ideal in $`L^1(𝔾)^{}`$, we see that $`gL^1(𝔾)`$. ∎ ###### Lemma 3.7. Let $`𝔾`$ be a locally compact quantum group, and let $`f,gL^1(G)`$ be states such that $`fg=gf=g`$. Then, if $`hL^1(𝔾)`$ is such that $`0hf`$, then $`hg=h(1)g`$ holds. ###### Proof. The proof of \[M–vD, Lemma 4.3\] carries over verbatim. ∎ We can now state and prove the first main result of this paper: ###### Theorem 3.8. The following are equivalent for a locally compact quantum group $`𝔾`$: 1. $`𝔾`$ is compact; 2. $`L^1(𝔾)`$ is an ideal in $`L^1(𝔾)^{}`$. ###### Proof. (i) $``$ (ii) is Corollary 3.5. (ii) $``$ (i): For each positive $`fL^1(𝔾)`$, set $$K_f:=\{gL^1(𝔾):g\text{ is a state such that }fg=f(1)g\}.$$ By Lemma 3.6, $`K_f`$ is non-empty, and since multiplication with $`f`$ from the left is weakly compact, we see that $`K_f`$ is a weakly compact subset of $`L^1(𝔾)`$ whenever $`f0`$. Let $`h,fL^1(𝔾)`$ be states with $`0hf`$. Then Lemma 3.7 yields that $`K_fK_h`$. As a consequence, we have $`K_{f_1+f_2}K_{f_1}K_{f_2}`$ for any positive $`f_1,f_2L^1(𝔾)`$. Fix a non-zero state $`f_0L^1(𝔾)`$. Then $`K_fK_{f_0}`$ is weakly compact for each positive $`fL^1(𝔾)`$ and non-empty (because it contains $`K_{f+f_0}`$). Consequently, $`\{K_f:fL^1(𝔾)\text{ is positive}\}`$ is non-empty. Any element of this intersection is a normal, left invariant state, so that $`𝔾`$ is compact. ∎ Let $`G`$ be a locally compact group, and let $`𝔾=(L^{\mathrm{}}(G),\mathrm{\Gamma }_G)`$. Since $`𝔾`$ is compact if and only $`G`$ is compact, the main result of \[Wat\] is a particular case of Theorem 3.8. On the other hand, $`\widehat{𝔾}=(\mathrm{VN}(G),\widehat{\mathrm{\Gamma }}_G)`$ is compact if and only if $`G`$ is discrete. From Theorem 3.8, we thus recover \[Lau, Theorem 3.7\]: ###### Corollary 3.9. Let $`G`$ be a locally compact group. Then $`A(G)`$ is an ideal in $`A(G)^{}`$ if and only if $`G`$ is discrete. ## 4 Discrete quantum groups There are various ways to define discrete quantum groups. Following \[Tom\], we say that the locally compact quantum group $`𝔾`$ is *discrete* if $`\widehat{𝔾}`$ is compact. As for compactness, we collect a few equivalent characterizations: ###### Proposition 4.1. The following are equivalent for a locally compact quantum group $`𝔾`$: 1. $`𝔾`$ is discrete; 2. there is a central minimal projection in $`𝒞_0(𝔾)`$; 3. $`L^1(𝔾)`$ has an identity. ###### Proof. (i) $``$ (ii) is noted in \[Tom, Remarks 3.8(2)\]. (ii) $``$ (i): Let $`p𝒞_0(𝔾)`$ be a central minimal projection. Then $`p`$ is also central and minimal in $`L^{\mathrm{}}(𝔾)`$. Let $`ϵL^1(𝔾)`$ denote the corresponding character. From the definition of $`\lambda `$, it is clear that $`\lambda (ϵ)𝒞_0(\widehat{𝔾})`$ is a unitary operator on $`L^2(\widehat{𝔾})`$. Consequently, $`𝒞_0(\widehat{𝔾})`$ has an identity, so that $`\widehat{𝔾}`$ is compact. (i) $``$ (iii): In \[Tom, Remarks 3.8(2)\], it is observed that the character $`ϵL^1(𝔾)`$—as in (ii) $``$ (i)—can also be chosen to satisfy $`(ϵ\mathrm{id})\mathrm{\Gamma }=(\mathrm{id}ϵ)\mathrm{\Gamma }`$ and thus is an identity for $`L^1(𝔾)`$. (iii) $``$ (i): If $`L^1(𝔾)`$ has an identity, then so has $`𝒞_0(\widehat{𝔾})=\overline{\lambda (L^1(𝔾))}`$, so that $`\widehat{𝔾}`$ is compact. ∎ As a trivial consequence of Theorem 3.8, a locally compact quantum group $`𝔾`$ is discrete if and only if $`L^1(\widehat{𝔾})`$ is an ideal in $`L^1(\widehat{𝔾})^{}`$. In this section, we give another characterization of discrete quantum groups in terms of second duals. For the proof, we require a general fact on Banach -algebras, which may be of independent interest. We say that a -representation $`\pi `$ of a Banach -algebra $`𝔄`$ on a Hilbert space $``$ is *non-degenerate* if the closed linear span of $`\{\pi (a)\xi :a𝔄,\xi \}`$ is all of $``$. In our next lemma, let $`𝒦()`$ denote the compact operators on a Hilbert space $``$. ###### Lemma 4.2. Let $`𝔄`$ be a Banach -algebra, let $``$ be a Hilbert space, and let $`\pi :𝔄()`$ be a faithful, non-degenerate -representation of $`𝔄`$ such that $`\pi (𝔄)𝒦()`$. Then, for each self-adjoint $`a𝔄`$, the spectra of $`a`$ in $`𝔄`$ and of $`\pi (a)`$ in $`()`$ coincide. ###### Proof. The claim is straightforward if $`dim<\mathrm{}`$. We thus limit ourselves to the case where $`dim=\mathrm{}`$. Let $`a𝔄`$ be self-adjoint, and let $`\mathrm{Sp}(a)`$ and $`\mathrm{Sp}(\pi (a))`$ denote the spectra of $`a`$ and $`\pi (a)`$ (in $`𝔄`$ and $`()`$, respectively). Since $`\pi (a)𝒦()`$ and since $`dim=\mathrm{}`$, we have $`0\mathrm{Sp}(\pi (a))`$. Also, as $`𝔄`$ cannot have an identity (again because $`\pi (𝔄)𝒦()`$), we have $`0\mathrm{Sp}(a)`$ as well. From \[Dal, Proposition 1.5.28\], we conclude that $`\mathrm{Sp}(\pi (a))\mathrm{Sp}(a)`$. For the converse inclusion, let $`\lambda \mathrm{Sp}(\pi (a))`$, and suppose without loss of generality that $`\lambda 0`$. From the spectral theory of compact operators on Hilbert space it is then well known that $`\lambda `$ is an eigenvalue of $`\pi (a)`$, i.e., there is $`\xi \{0\}`$ such that $`(\lambda \pi (a))\xi =0`$. Let $`𝔄^\mathrm{\#}`$ denote the *unitization* of $`𝔄`$, i.e., the algebra $`𝔄`$ with an identity adjoined, and assume that $`\lambda \mathrm{Sp}(a)`$. By the definition of $`\mathrm{Sp}(a)`$ (see \[Dal, p. 78\], for instance), there is $`x𝔄^\mathrm{\#}`$ such that $`x(\lambda a)=1`$. The representation $`\pi `$ extends canonically to a unital -representation $`\pi ^\mathrm{\#}:𝔄^\mathrm{\#}()`$, so that we obtain $$\xi =\pi ^\mathrm{\#}(1)\xi =\pi ^\mathrm{\#}(x)(\lambda \pi (a))\xi =0,$$ which is a contradiction. ∎ By a $`C^{}`$-norm on a Banach -algebra $`𝔄`$ we mean a submultiplicative norm $`\gamma `$ on $`𝔄`$ satisfying $`\gamma (a^{}a)=\gamma (a)^2`$ for $`a𝔄`$. We denote the completion of $`𝔄`$ with respect to $`\gamma `$ by $`C_\gamma ^{}(𝔄)`$, which is a $`C^{}`$-algebra. We say that $`𝔄`$ has a *unique $`C^{}`$-norm* if there is only one $`C^{}`$-norm on $`𝔄`$. We also require the notion of a regular Banach algebra (\[Dal, Definition 4.1.16\]): a commutative Banach algebra $`𝔄`$ with charater space $`\mathrm{\Phi }_𝔄`$ is called *regular* if, for each closed $`F\mathrm{\Phi }_𝔄`$ and $`\varphi \mathrm{\Phi }_𝔄F`$, there is $`a𝔄`$ with $`\widehat{a}|_F0`$ and $`\widehat{a}(\varphi )=1`$, where $`\widehat{a}`$ stands for the Gelfand transform of $`a`$. ###### Proposition 4.3. Let $`𝔄`$ be a Banach -algebra, and suppose that there is a $`C^{}`$-norm $`\gamma `$ on $`𝔄`$ such that $`C_\gamma ^{}(𝔄)`$ is an ideal in $`C_\gamma ^{}(𝔄)^{}`$. Then $`𝔄`$ has a unique $`C^{}`$-norm. ###### Proof. Suppose without loss of generality that $`𝔄`$ is infinite-dimensional. Let $`a𝔄`$ be self-adjoint, and let $`a`$ denote the closed subalgebra of $`𝔄`$ generated by $`a`$. We claim that the Banach algebra $`a`$ is regular. Since $`C_\gamma ^{}(𝔄)`$ is an ideal in $`C_\gamma ^{}(𝔄)^{}`$, it follows from \[Tak 1, Exercises III.5.3 and III.5.4\], that there is a family $`(_\alpha )_\alpha `$ of Hilbert spaces such that $$C_\gamma ^{}(𝔄)c_0\text{-}\underset{\alpha }{}𝒦(_\alpha )𝒦(),$$ where $`:=\mathrm{}^2\text{-}_\alpha _\alpha `$. This yields a faithful, non-degenerate -representation $`\pi :𝔄()`$ with $`\pi (𝔄)𝒦()`$. From Lemma 4.2 and the spectral theory of compact operators on Hilbert space we conclude that the spectrum of $`a`$ in $`𝔄`$ is of the form $`\{\lambda _n:n\}\{0\}`$, where $`\lambda _n0`$ for $`n`$ and $`lim_n\mathrm{}\lambda _n=0`$. Consequently, the character space of $`a`$ is canonically homeomorphic to the discrete subset $`\{\lambda _n:n\}`$ of $``$. With the help of the holomorphic functional calculus, it is then straightforward to see that $`a`$ is regular, as claimed. Since $`a𝔄`$ was an arbitrary self-adjoint element, $`𝔄`$ is therefore *locally regular* in the sense of \[Bar, Definition 4.1\] and thus has a unique $`C^{}`$-norm (\[Bar, Lemma 4.2\]). ∎ ###### Theorem 4.4. The following are equivalent for a locally compact quantum group $`𝔾`$: 1. $`𝔾`$ is discrete; 2. $`𝒞_0(𝔾)`$ is an ideal in $`𝒞_0(𝔾)^{}`$. ###### Proof. (i) $``$ (ii): Let $$I:=\{a𝒞_0(𝔾):𝒞_0(𝔾)bba\text{ is weakly compact}\},$$ and note that $`I`$ is closed. Since $`𝔾`$ is discrete, $`\widehat{𝔾}`$ is compact, so that we have the map $`\iota :L^2(\widehat{𝔾})L^1(\widehat{𝔾})`$ as defined immediately prior to Lemma 3.3; moreover, we have $$\widehat{\lambda }(f)\widehat{\lambda }(\iota (\xi ))=\widehat{\lambda }(f\iota (\xi ))=\widehat{\lambda }(\iota (\widehat{\lambda }(f)\xi ))(fL^1(\widehat{𝔾}),\xi L^2(\widehat{𝔾})).$$ Since $`L^2(\widehat{𝔾})`$ is reflexive, it follows—as in the proof of Proposition 3.4—that $`\widehat{\lambda }(\iota (L^2(\widehat{𝔾})))`$ is contained in $`I`$. Since $`\widehat{\lambda }(\iota (L^2(\widehat{𝔾})))`$ is dense in $`𝒞_0(𝔾)`$, $`I`$ must equal all of $`𝒞_0(𝔾)`$, i.e., right multiplication in $`𝒞_0(𝔾)`$ is weakly compact. Using the involution on $`𝒞_0(𝔾)`$, we see that left multiplication in $`𝒞_0(𝔾)`$ is also weakly compact, so that $`𝒞_0(𝔾)`$ is an ideal in $`𝒞_0(𝔾)^{}`$. (ii) $``$ (i): Let $`\gamma `$ denote the $`C^{}`$-norm on $`L_{}^1(\widehat{𝔾})`$ induced by $`\widehat{\lambda }`$, so that $`C_\gamma ^{}(L_{}^1(\widehat{𝔾}))=𝒞_0(𝔾)`$. From Proposition 4.3, we conclude that $`L_{}^1(\widehat{𝔾})`$ has a unique $`C^{}`$-norm; in particular, $`𝒞_0^u(𝔾)𝒞_0(𝔾)`$ holds. Consequently, $`𝒞_0(𝔾)`$ has a character, say $`ϵ`$, by \[B–T, Theorem 3.1\]. Let $`p𝒞_0(𝔾)^{}`$ be the corresponding central minimal projection. Choose $`a𝒞_0(𝔾)`$ with $`ϵ(a)=1`$. It follows that $`p=ϵ(a)p=ap𝒞_0(𝔾)`$. By Proposition 4.1, this means that $`𝔾`$ is discrete. ∎ Since compactness and discreteness are dual to each other, we finally obtain as a consequence of both Theorems 3.8 and 4.4: ###### Corollary 4.5. The following are equivalent for a locally compact quantum group $`𝔾`$: 1. $`𝔾`$ is compact; 2. $`L^1(𝔾)`$ is an ideal in $`L^1(𝔾)^{}`$; 3. $`𝒞_0(\widehat{𝔾})`$ is an ideal in $`𝒞_0(\widehat{𝔾})^{}`$; 4. $`\widehat{𝔾}`$ is discrete. \[\] | Author’s address: | Department of Mathematical and Statistical Sciences | | --- | --- | | | University of Alberta | | | Edmonton, Alberta | | | Canada T6G 2G1 | | E-mail: | vrunde@ualberta.ca | | URL: | http://www.math.ualberta.ca/runde/ |
warning/0506/hep-ph0506322.html
ar5iv
text
# 𝐵_𝑑⁢(𝐵̄_𝑑)→{𝜌^±⁢𝜋^∓,𝜌⁺⁢𝜌⁻,𝜋⁺⁢𝜋⁻}: hunting for alpha ## 1 Introduction In paper from the data on CP asymmetries in $`B_d(\overline{B}_d)\rho ^\pm \pi ^{}`$, $`\rho ^+\rho ^{}`$ decays and BABAR data on CP asymmetries in $`B_d(\overline{B}_d)\pi ^+\pi ^{}`$ decays we determine the value of angle $`\alpha `$ of the unitarity triangle: $$\alpha =96^o\pm 3^o,$$ (1) where only a tree quark decay amplitude $`\overline{b}u\overline{u}\overline{d}(bu\overline{u}d)`$ was taken into account. The numerical values of angle $`\alpha `$ obtained from the considered decays are consistent with each other and with the value which follows from the global CKM fit. This observation testifies to the validity of a proposed approach. As the next step in the present paper we will study what changes in the values of $`\alpha `$ are induced by QCD penguins. Our aim is twofold. First, in this way we will get an estimate of the theoretical uncertainty of the value of $`\alpha `$ determined in paper . Secondly, we will get the formulas for CP violating parameters describing these decays which vanish when penguins are neglected ($`C_{\rho \pi }`$, $`A_{CP}^{\rho \pi }`$, $`C_{\rho \rho }`$, $`C_{\pi \pi }`$). The angle shifts $`\mathrm{\Delta }\alpha `$ we are interested in were estimated in paper ; however, in that paper FSI phases were neglected, that is why $`C_{\rho \pi }=A_{CP}^{\rho \pi }=C_{\rho \rho }=C_{\pi \pi }=0`$ follows from (the nonzero penguin amplitudes are a necessary, but not a sufficient condition for $`C_{\rho \pi }\mathrm{}`$ to be nonzero). We will take these phases into account. The asymmetries depend on the differences of FSI phases in the processes described by the tree and penguin diagrams. One source of these differences is an imaginary part of the quark penguin diagram, the so-called BSS mechanism of the strong phases generation (see also paper ). The phase of the penguin diagram depends on the gluon $`q^2`$ which is transferred to $`u\overline{u}`$ pair, each quark of which goes to different $`\pi ^\pm `$\- or $`\rho ^\pm `$-mesons. In this way the value of $`q^2`$ depends on the light meson wave functions and we can estimate it only roughly. Another source of FSI phases is hadron rescattering and even less is known about the values of the phase shifts between the penguin and tree diagrams generated in this way. In view of this we will determine FSI phases from the experimental data on CPV asymmetries, and investigate to what values of $`\alpha `$ it will lead. In Appendix we present the weak interaction Hamiltonian which is responsible for $`bu\overline{u}d`$ transition and calculate the necessary matrix elements. Using these formulas in sections 2, 3, and 4 we study $`B\rho \pi `$, $`\rho \rho `$ and $`\pi \pi `$ decays correspondingly and extract the values of angle $`\alpha `$ from the experimental data on CP asymmetries in these decays. We conclude in section 5 with the averaged value of $`\alpha `$ and a general discussion. ## 2 $`\alpha `$ from $`\overline{B}_d(B_d)\rho ^{}\pi ^\pm `$ The time dependence of the decay probabilities is given by : $`{\displaystyle \frac{dN(B_d(\overline{B}_d)\rho ^\pm \pi ^{})}{dt}}`$ $`=`$ $`(1\pm A_{CP}^{\rho \pi })e^{t/\tau }[1q(C_{\rho \pi }\pm \mathrm{\Delta }C_{\rho \pi })\times `$ (2) $`\times `$ $`\mathrm{cos}(\mathrm{\Delta }mt)+q(S_{\rho \pi }\pm \mathrm{\Delta }S_{\rho \pi })\mathrm{sin}(\mathrm{\Delta }mt)],`$ where $`q=1`$ describes the case when at $`t=0`$ $`B_d`$ was produced, while $`q=1`$ corresponds to $`\overline{B}_d`$ production at $`t=0`$. In the case of $`\mathrm{{\rm Y}}(4S)B_d\overline{B}_d`$ decay the flavor of the beauty meson which will decay to $`\rho \pi `$ is tagged by the charge of a lepton in the other beauty meson semileptonic decay. A partner decay starts clocks as well. $`\tau `$ is $`B_d(\overline{B}_d)`$ life time, while $`\mathrm{\Delta }m`$ is the difference of masses of $`(B_d,\overline{B}_d)`$ system eigenstates (it equals the frequency of $`B_d\overline{B}_d`$ oscillations). From Eqs. (A15) and (A17) we obtain: $$\overline{M}^+=AV_{ub}V_{ud}^{}[10.07e^{i(\delta \alpha )}]=AV_{ub}V_{ud}^{}[10.07\mathrm{sin}\delta +i0.07\mathrm{cos}\delta ],$$ (3) $$M^+=BV_{ub}^{}V_{ud},$$ (4) $$\lambda ^+\frac{q}{p}\frac{\overline{M}^+}{M^+}=e^{2i\alpha }\frac{A}{B}[10.07\mathrm{sin}\delta +i0.07\mathrm{cos}\delta ],$$ (5) where parameters $`q`$ and $`p`$ enter the expressions for $`(B_d,\overline{B}_d)`$ eigenstates and we have substituted $`\alpha =\pi /2`$ in the (small) second term in square brackets in Eq. (A15). Analogously we get: $$\overline{M}^+=BV_{ub}V_{ud}^{},$$ (6) $$M^+=AV_{ub}^{}V_{ud}[10.07e^{i(\alpha +\delta )}]=AV_{ub}^{}V_{ud}[1+0.07\mathrm{sin}\delta 0.07i\mathrm{cos}\delta ],$$ (7) $$\lambda ^+\frac{q}{p}\frac{\overline{M}^+}{M^+}=e^{2i\alpha }\frac{B}{A}[1+0.07\mathrm{sin}\delta i0.07\mathrm{cos}\delta ]^1.$$ (8) From the expressions for the quantities $`C_{\rho \pi }`$ and $`\mathrm{\Delta }C_{\rho \pi }`$ : $$C_{\rho \pi }\pm \mathrm{\Delta }C_{\rho \pi }=\frac{1|\lambda ^\pm |^2}{1+|\lambda ^\pm |^2}$$ (9) we obtain: $`\mathrm{\Delta }C_{\rho \pi }`$ $`=`$ $`{\displaystyle \frac{a^2b^2}{a^2+b^2}},{\displaystyle \frac{a^2}{b^2}}={\displaystyle \frac{1+\mathrm{\Delta }C_{\rho \pi }}{1\mathrm{\Delta }C_{\rho \pi }}},`$ $`C_{\rho \pi }`$ $`=`$ $`0.28\mathrm{sin}\delta {\displaystyle \frac{(a/b)^2}{(1+a^2/b^2)^2}},`$ (10) where $`a|A|`$, $`b|B|`$. The Belle and BABAR averaged result for $`\mathrm{\Delta }C_{\rho \pi }`$ is : $$\mathrm{\Delta }C_{\rho \pi }=0.22\pm 0.10,$$ (11) which leads to: $$\left(\frac{a}{b}\right)^2=1.56\pm 0.33.$$ (12) From the averaged experimental result $$C_{\rho \pi }=0.31\pm 0.10$$ (13) and Eq. (10) we get: $$\mathrm{sin}\delta =4.6\pm 1.5.$$ (14) We see that poor accuracy in the measurement of $`C_{\rho \pi }`$ does not allow to get any definite information on the value of phase $`\delta `$. The next observable we wish to discuss is CP asymmetry $`A_{CP}^{\rho \pi }`$: $`A_{CP}^{\rho \pi }`$ $`=`$ $`{\displaystyle \frac{|M^+|^2|\overline{M}^+|^2+|\overline{M}^+|^2|M^+|^2}{|M^+|^2+|\overline{M}^+|^2+|\overline{M}^+|^2+|M^+|^2}}=`$ (15) $`=`$ $`0.14\mathrm{sin}\delta {\displaystyle \frac{(a/b)^2}{1+(a/b)^2}},`$ which should be compared with the experimental result : $$A_{CP}^{\rho \pi }=0.102\pm 0.045.$$ (16) From (15) and (16) we get: $$\mathrm{sin}\delta =1.2\pm 0.5,$$ (17) and it differs from given in Eq.(14) by 3.5 standard deviations. This is the largest discrepancy we encounter in this paper. Averaging these two numbers we obtain: $$\mathrm{sin}\delta =0.62\pm 0.47.$$ (18) Finally we come to the discussion of the observables which are sensitive to the angle $`\alpha `$: $$S_{\rho \pi }\pm \mathrm{\Delta }S_{\rho \pi }=\frac{2\mathrm{I}\mathrm{m}\lambda ^\pm }{1+|\lambda ^\pm |^2},$$ (19) $$S_{\rho \pi }=\frac{2a/b}{1+a^2/b^2}[\mathrm{sin}2\alpha \mathrm{cos}\stackrel{~}{\delta }0.07\mathrm{cos}\delta \mathrm{cos}\stackrel{~}{\delta }0.07\frac{a^2/b^21}{a^2/b^2+1}\mathrm{sin}\delta \mathrm{sin}\stackrel{~}{\delta }],$$ (20) $`\mathrm{\Delta }S_{\rho \pi }`$ $`=`$ $`{\displaystyle \frac{2a/b}{1+a^2/b^2}}[\mathrm{cos}2\alpha \mathrm{sin}\stackrel{~}{\delta }+0.07\mathrm{cos}\delta \mathrm{sin}\stackrel{~}{\delta }\mathrm{sin}2\alpha `$ (21) $``$ $`0.07{\displaystyle \frac{a^2/b^21}{a^2/b^2+1}}\mathrm{sin}\delta \mathrm{cos}\stackrel{~}{\delta }\mathrm{sin}2\alpha ],`$ where the definition of the phase $`\stackrel{~}{\delta }`$ is $`A/B(a/b)e^{i\stackrel{~}{\delta }}`$ and in the small terms proportional to 0.07 in the expression for $`S_{\rho \pi }`$ we have substituted $`\mathrm{cos}2\alpha =1`$. Let us start the analysis of the experimental data from $`\mathrm{\Delta }S_{\rho \pi }`$. According to : $$\mathrm{\Delta }S_{\rho \pi }=0.09\pm 0.13,$$ (22) which is much less than one. According to Eq. (12) the factor which multiplies square brackets in Eq. (20) (and in Eq. (21)) is very close to one, that is why it is the expression in square brackets which should be much less than one. The second and the third terms of this expression are really very small and we can neglect them. What concerns the first term, it is small when $`\stackrel{~}{\delta }`$ is close to zero or $`\pi `$: $$\mathrm{sin}\stackrel{~}{\delta }=\mathrm{\Delta }S_{\rho \pi }=0.09\pm 0.13,$$ (23) where a small deviation of $`\mathrm{cos}2\alpha `$ from – 1 is neglected. Now everything is ready and from Eq. (20) we get: $$\mathrm{sin}2\alpha =S_{\rho \pi }/\mathrm{cos}\stackrel{~}{\delta }+0.07\mathrm{cos}\delta ,$$ (24) where we omit the last term in square brackets since it is negligibly small. In order to find angle $`\alpha `$ from the experimental data : $$S_{\rho \pi }=0.13\pm 0.13,$$ (25) we should determine the values of $`\mathrm{cos}\delta `$ and $`\mathrm{cos}\stackrel{~}{\delta }`$. To propagate errors from $`\mathrm{sin}\delta `$ to $`\mathrm{cos}\delta `$, we consider gaussian distribution for $`\mathrm{sin}\delta `$ truncated to physical region $`|\mathrm{sin}\delta |<1`$, transform it into (non-gaussian) distribution for $`\mathrm{cos}\delta `$ and take an interval containing 68% of probability. In this way from Eq. (18) we obtain: $$|\mathrm{cos}\delta |=0.88\pm 0.12.$$ (26) The average value of $`|\mathrm{cos}\delta |`$ appears to be close to one due to the so-called Jacobian pick. Concerning $`\stackrel{~}{\delta }`$ it follows from Eq. (23) that $`|\mathrm{cos}\stackrel{~}{\delta }|=1`$ with very good accuracy. Depending on the values of phases $`\stackrel{~}{\delta }`$ and $`\delta `$ we get the following four domains for the angle $`\alpha `$: $`\stackrel{~}{\delta }0,\delta 0:`$ $`\mathrm{sin}2\alpha =0.13\pm 0.13+0.06`$ (27) $`\alpha =92^o\pm 4^o`$ $`\stackrel{~}{\delta }0,\delta \pi :`$ $`\mathrm{sin}2\alpha =0.13\pm 0.130.06`$ (28) $`\alpha =96^o\pm 4^o`$ $`\stackrel{~}{\delta }\pi ,\delta 0:`$ $`\mathrm{sin}2\alpha =0.13\pm 0.13+0.06`$ (29) $`\alpha =84^o\pm 4^o`$ $`\stackrel{~}{\delta }\pi ,\delta \pi :`$ $`\mathrm{sin}2\alpha =0.13\pm 0.130.06`$ (30) $`\alpha =88^o\pm 4^o.`$ Thus QCD penguins split values of $`\alpha `$ obtained without taking them into account: $`94^o92^o,96^o`$; $`86^o84^o,88^o`$. If BSS mechanism is valid, then only the domains given by Eqs.(27) and (29) remain (see also ). ## 3 $`\alpha `$ from $`\overline{B}_d(B_d)\rho ^+\rho ^{}`$ The time dependence of CP violating asymmetry is described by the following formula: $$a_{CP}(t)\frac{\frac{dN(\overline{B}_d\rho _L^+\rho _L^{})}{dt}\frac{dN(B_d\rho _L^+\rho _L^{})}{dt}}{\frac{dN(\overline{B}_d\rho _L^+\rho _L^{})}{dt}+\frac{dN(B_d\rho _L^+\rho _L^{})}{dt}}=C_{\rho \rho }\mathrm{cos}(\mathrm{\Delta }mt)+S_{\rho \rho }\mathrm{sin}(\mathrm{\Delta }mt).$$ (31) Let us remind that the longitudinal polarization fraction $`f_L=0.98\pm 0.01\pm 0.02`$ ; $`f_L=0.95\pm 0.03\pm 0.03`$ and its closeness to one greatly simplify the extraction of CPV parameters from $`B_d(\overline{B}_d)\rho ^+\rho ^{}`$ decay data. These parameters are given by the following expressions: $$C_{\rho \rho }=\frac{1|\lambda _{\rho \rho }|^2}{1+|\lambda _{\rho \rho }|^2},S_{\rho \rho }=\frac{2\mathrm{I}\mathrm{m}\lambda _{\rho \rho }}{1+|\lambda _{\rho \rho }|^2}.$$ (32) From Eq. (A14) we obtain: $$\lambda _{\rho \rho }\frac{q}{p}\frac{\overline{M}_{\rho \rho }}{M_{\rho \rho }}=e^{2i\alpha }\frac{10.07e^{i(\delta \alpha )}}{10.07e^{i(\delta +\alpha )}}=e^{2i\alpha }(1+0.14i\mathrm{sin}\alpha e^{i\delta }),$$ (33) where we use the same letter $`\delta `$ for FSI phase difference of the amplitudes generated by penguin and tree diagrams as in the case of $`B\rho \pi `$ decays. These differences would be really the same if BSS mechanism dominates. Comparing the averaged experimental result for $`C_{\rho \rho }`$ $$C_{\rho \rho }=0.03\pm 0.18\pm 0.09\text{[8]},$$ (34) $$C_{\rho \rho }=0.0\pm 0.30\pm 0.10\text{[9]};$$ (35) $$C_{\rho \rho }^{exp}=0.02\pm 0.17$$ (36) with the theoretical expression which follows from Eqs. (32), (33) $$C_{\rho \rho }=0.14\mathrm{sin}\alpha \mathrm{sin}\delta 0.14\mathrm{sin}\delta $$ (37) we get: $$\mathrm{sin}\delta =0.15\pm 1.2;|\mathrm{cos}\delta |=0.88\pm 0.12,$$ (38) where the same procedure of error propagation as in the case of $`B\rho \pi `$ was used and we sum statistical and systematic errors of $`\mathrm{sin}\delta `$ as independent. From Eq. (32) we obtain: $$S_{\rho \rho }=\mathrm{sin}2\alpha +0.14\mathrm{sin}\alpha \mathrm{cos}2\alpha \mathrm{cos}\delta =\mathrm{sin}2\alpha 0.14\mathrm{cos}\delta .$$ (39) According to the recent measurements: $$S_{\rho \rho }=0.33\pm 0.26\text{[8]},$$ (40) $$S_{\rho \rho }=0.09\pm 0.43\text{[9]},$$ (41) $$S_{\rho \rho }^{exp}=0.21\pm 0.22,$$ (42) and we get two domains for $`\alpha `$: $$\delta 0:\alpha =92^o\pm 7^o$$ (43) $$\delta \pi :\alpha =100^o\pm 7^o.$$ (44) Just as in the case of $`B\rho \pi `$ decays only the first domain remains if $`|\delta |<\pi /2`$ . Let us note that using the isospin analysis (which allows to prove the smallness of the penguin contribution) it was obtained: $$\alpha =100^o\pm 13^o\text{[8]},$$ (45) $$\alpha =87^o\pm 17^o\text{[9]},$$ (46) where the error is mainly due to the uncertainty of the penguin contribution. Extracting this uncertainty and averaging last two numbers we get: $$\alpha =96^o\pm 7^o(exp)\pm 11^o(penguin).$$ (47) ## 4 $`\alpha `$ from $`\overline{B}_d(B_d)\pi ^+\pi ^{}`$ The time dependence of CP violating asymmetry is given by the formula analogous to Eq. (31) with the evident substitution of $`\pi `$ instead of $`\rho `$. Eq. (32) with the same substitution is valid as well, while for the quantity $`\lambda _{\pi \pi }`$ from Eq. (A13) we obtain: $$\lambda _{\pi \pi }\frac{q}{p}\frac{\overline{M}_{\pi ^+\pi ^{}}}{M_{\pi ^+\pi ^{}}}=e^{2i\alpha }(1+0.28i\mathrm{sin}\alpha e^{i\delta }),$$ (48) and what concerns letter $`\delta `$ we should repeat the comment made after Eq. (33). The experimental data of BABAR and Belle for CPV parameters $`S_{\pi \pi }`$ and $`C_{\pi \pi }`$ were controversial though at present (with the latest Belle results) the divergence diminishes. In view of this we will perform a two step analysis, taking at the beginning only BABAR results and then the averaged results of two collaborations. Comparing the theoretical expression $$C_{\pi \pi }=0.28\mathrm{sin}\alpha \mathrm{sin}\delta 0.28\mathrm{sin}\delta $$ (49) with BABAR result $$C_{\pi \pi }^{\mathrm{BABAR}}=0.09\pm 0.15$$ (50) we get: $$\mathrm{sin}\delta =0.32\pm 0.54,$$ (51) while for $`S_{\pi \pi }`$ we have: $$S_{\pi \pi }=\mathrm{sin}2\alpha 0.28\mathrm{cos}\delta ,S_{\pi \pi }^{\mathrm{BABAR}}=0.30\pm 0.17.$$ (52) From Eq. (51) we get: $$|\mathrm{cos}\delta |=0.9\pm 0.1$$ (53) and two domains of $`\alpha `$ corresponding to two signs of $`\mathrm{cos}\delta `$<sup>1</sup><sup>1</sup>1Assuming $`|\delta |<\pi /2`$ we would get only the first domain.: $$\delta 0:\alpha =91^o\pm 5^o(\mathrm{exp})\pm 1^o(\mathrm{theor})$$ (54) $$\delta \pi :\alpha =107^o\pm 5^o(\mathrm{exp})\pm 1^o(\mathrm{theor}).$$ (55) Belle result: $$C_{\pi \pi }^{\mathrm{Belle}}=0.56\pm 0.13$$ (56) deviates by $`2.5\sigma `$ from BABAR and (if correct) would require considerably larger $`P/T`$ ratio than we use in our paper. Finally, averaging (50) and (56) one gets : $$C_{\pi \pi }^{\mathrm{ex}}=0.37\pm 0.10,$$ (57) and comparing with the theoretical expression (49) we obtain: $$\mathrm{sin}\delta =1.32\pm 0.35,$$ (58) which leads to: $$|\mathrm{cos}\delta |=0.55\pm 0.25.$$ (59) From the averaged experimental result : $$S_{\pi \pi }^{\mathrm{ex}}=0.50\pm 0.12(\mathrm{exp})$$ (60) we get two domains<sup>2</sup><sup>2</sup>2Assuming $`|\delta |<\pi /2`$ which follows from BSS mechanism of $`\delta `$ generation, we would get the first domain only.: $$\delta 0:\alpha =100^o\pm 4^o(\mathrm{exp})\pm 2^o(\mathrm{theor})$$ (61) $$\delta \pi :\alpha =110^o\pm 4^o(\mathrm{exp})\pm 2^o(\mathrm{theor}).$$ (62) ## 5 Conclusions We have analyzed CPV asymmetries in $`B_d(\overline{B}_d)\rho ^\pm \pi ^{}`$, $`\rho _L^+\rho _L^{}`$ and $`\pi ^+\pi ^{}`$ decays induced by the charmless strangeless $`b`$-quark decay $`bu\overline{u}d`$. This decay can proceed through a tree or penguin diagram. As it was noted in when the penguin diagram is neglected, one obtains the values of the unitarity triangle angle $`\alpha `$ from CPV asymmetries in these decays which are consistent with each other as well as with the value of $`\alpha `$ which follows from the global CKM fit. However, in order to determine the theoretical accuracy of $`\alpha `$ extracted from the decays under study one should take the penguin amplitude into account. This was done in the present paper, where the moduli of penguin over tree ratios were calculated with the help of the factorization: $$<M_1M_2|j_1j_2|B>=<M_1|j_1|B><M_2|j_2|0>,$$ (63) while FSI phase shifts between the tree and penguin amplitudes were extracted from experimental data. In order to determine numerical value of $`\alpha `$ one should average the values which follow from the considered decays. Since the phase shifts $`\delta `$ can be different in $`B\rho \rho ,\pi \pi `$ and $`\rho \pi `$ decays, we get too many possibilities. That is why let us limit ourselves to the theoretically motivated case $`|\delta |<\pi /2`$. Averaging Eqs. (43) and (61) we obtain: $$\delta 0:\alpha _{\rho \rho ,\pi \pi }=98^o\pm 4^o.$$ (64) Averaging it with (27) and (29) we get two possibilities: $`\alpha _{bu\overline{u}d}`$ $`=`$ $`95^o\pm 3^o,\mathrm{or}`$ $`\alpha _{bu\overline{u}d}`$ $`=`$ $`91^o\pm 3^o,`$ (65) and the last one corresponds to the smallest possible value of $`\alpha `$. In the case $`|\delta _i|>\pi /2`$, $`\stackrel{~}{\delta }=0`$ averaging Eqs.(28), (44), (62) we get the largest possible value: $`\alpha _{bu\overline{u}d}=102^o\pm 3^o`$. The global fit results for $`\alpha `$ are: $$\alpha _{\mathrm{UTfit}}^{\text{[10]}}=94^o\pm 8^o,\alpha _{\mathrm{CKMfitter}}^{\text{[11]}}=94\pm 10^o.$$ (66) Thus the accuracy of the present day knowledge of $`\alpha `$ can be close to that of $`\beta `$: $$\beta =23^o\pm 2^o.$$ (67) I am grateful to A.V. Fedotov for great help in the treatment of the experimental data. I would like to thank A.E. Bondar and M.B. Voloshin for useful remarks. This work was partially supported by grants NSh-2328.2003.2 and RFBR 05-02-17203. Appendix The strong interaction renormalization of the tree Hamiltonian which describes the beauty hadrons weak decays is much smaller than for the case of the strange particle decays since the masses of beauty hadrons are much closer to $`M_W`$ in the logarithmic scale. In the leading logarithmic approximation for operators $`O_1`$ and $`O_2`$ we have: $`\widehat{H}_{1,2}`$ $`=`$ $`{\displaystyle \frac{G_F}{\sqrt{2}}}V_{ub}V_{ud}^{}\{\left[{\displaystyle \frac{\alpha _S(m_b)}{\alpha _S(M_W)}}\right]^{4/b}[\overline{u}\gamma _\alpha (1+\gamma _5)b\overline{d}\gamma _\alpha /1+\gamma _5)u`$ (A1) $``$ $`\overline{d}\gamma _\alpha (1+\gamma _5)b\overline{u}\gamma _\alpha (1+\gamma _5)u]+[{\displaystyle \frac{\alpha _S(m_b)}{\alpha _S(M_W)}}]^{2/b}[\overline{u}\gamma _\alpha (1+\gamma _5)b\times `$ $`\times `$ $`\overline{d}\gamma _\alpha (1+\gamma _5)u+\overline{d}\gamma _\alpha (1+\gamma _5)b\overline{u}\gamma _\alpha (1+\gamma _5)u]\}`$ and substituting $`\alpha _S(M_W)=0.12`$, $`\alpha _S(m_b)=0.2`$, $`b=23/3`$ we get: $`\widehat{H}_{1,2}`$ $`=`$ $`{\displaystyle \frac{G_F}{\sqrt{2}}}V_{ub}V_{ud}^{}\{1.1\overline{u}\gamma _\alpha (1+\gamma _5)b\overline{d}\gamma _\alpha (1+\gamma _5)u`$ (A2) $``$ $`0.2\overline{d}\gamma _\alpha (1+\gamma _5)b\overline{u}\gamma _\alpha (1+\gamma _5)u\}.`$ NLO calculations confirm and refine this result. From Table 1 of for the value $`\mathrm{\Lambda }_4=280`$ MeV (which corresponds to $`\alpha _S(M_Z)=0.118`$) we get 1.14 instead of our 1.1 and -0.31 instead of our -0.2. At one loop the following QCD penguin operator is generated: $`\widehat{H}_{36}`$ $`=`$ $`{\displaystyle \frac{G_F}{\sqrt{2}}}(V_{cb}V_{cd}^{}+V_{ub}V_{ud}^{}){\displaystyle \frac{\alpha _S(m_b)}{12\pi }}\mathrm{ln}\left({\displaystyle \frac{M_W}{m_b}}\right)^2(\overline{d}\gamma _\mu (1+\gamma _5)\stackrel{}{\lambda }b)\times `$ (A3) $`\times `$ $`(\overline{u}\gamma _\mu \stackrel{}{\lambda }u+\overline{d}\gamma _\mu \stackrel{}{\lambda }d)=+{\displaystyle \frac{G_F}{\sqrt{2}}}V_{tb}V_{td}^{}0.03\{{\displaystyle \frac{2}{3}}\overline{d}\gamma _\alpha (1+\gamma _5)b\times `$ $`\times `$ $`(\overline{u}\gamma _\alpha u+\overline{d}\gamma _\alpha d)+2(\overline{d}_a\gamma _\alpha (1+\gamma _5)b^c)(\overline{u}_c\gamma _\alpha u^\alpha +\overline{d}_c\gamma _\alpha d^a)\},`$ where $`\stackrel{}{\lambda }`$ are eight colour SU(3) Gell-Mann matrices, and Fierz identity $`\stackrel{}{\lambda }_{ab}\stackrel{}{\lambda }_{cd}=2/3\delta _{ab}\delta _{cd}+2\delta _{ad}\delta _{bc}`$ as well as unitarity relation $`V_{cb}V_{cd}^{}+V_{ub}V_{ud}^{}=V_{tb}V_{td}^{}`$ were used. Substituting $`\overline{q}\gamma _\alpha q=\frac{1}{2}\overline{q}\gamma _\alpha (1+\gamma _5)q+\frac{1}{2}\overline{q}\gamma _\alpha (1\gamma _5)q`$ we find the renormalization factors +0.01 and -0.03 for operators $`O_3`$, $`O_5`$ and $`O_4`$, $`O_6`$ respectively. At NLO the renormalization factors for the operators in which only the left-handed quarks are involved ($`O_3,O_4`$) are different from those for the operators in which both left- and right-handed quarks participate ($`O_5,O_6`$). From the same Table 1 of we get 0.016 and 0.010 instead of 0.01 and -0.036 and -0.045 instead of -0.03. Finally, the effective Hamiltonian which describes the charmless strangeless $`\overline{B}_d`$ decays looks like: $$\widehat{H}=\frac{G_F}{\sqrt{2}}\left[V_{ub}V_{ud}^{}(c_1O_1+c_2O_2)V_{tb}V_{td}^{}(c_3O_3+c_4O_4+c_5O_5+c_6O_6)\right],$$ (A4) $$\begin{array}{cc}O_1=\overline{u}\gamma _\alpha (1+\gamma _5)b\overline{d}\gamma _\alpha (1+\gamma _5)u\hfill & c_1=1.14,\hfill \\ O_2=\overline{d}\gamma _\alpha (1+\gamma _5)b\overline{u}\gamma _\alpha (1+\gamma _5)u\hfill & c_2=0.31,\hfill \\ O_3=\overline{d}\gamma _\alpha (1+\gamma _5)b[\overline{u}\gamma _\alpha (1+\gamma _5)u+\overline{d}\gamma _\alpha (1+\gamma _5)d]\hfill & c_3=0.016,\hfill \\ O_4=\overline{d}_a\gamma _\alpha (1+\gamma _5)b^c[\overline{u}_c\gamma _\alpha (1+\gamma _5)u^a+\overline{d}_c\gamma _\alpha (1+\gamma _5)d^a]\hfill & c_4=0.036,\hfill \\ O_5=\overline{d}\gamma _\alpha (1+\gamma _5)b[\overline{u}\gamma _\alpha (1\gamma _5)u+\overline{d}\gamma _\alpha (1\gamma _5)d]\hfill & c_5=0.010,\hfill \\ O_6=\overline{d}_a\gamma _\alpha (1+\gamma _5)b^c[\overline{u}_c\gamma _\alpha (1\gamma _5)u^a+\overline{d}_c\gamma _\alpha (1\gamma _5)d^a]\hfill & c_6=0.045,\hfill \end{array}$$ (A5) and the complex conjugate Hamiltonian describes $`B_d`$ decays. Our next task is to calculate the matrix elements of $`\widehat{H}`$ between $`\overline{B}_d`$ and $`\rho ^\pm \pi ^{}`$, $`\rho ^+\rho ^{}`$ and $`\pi ^+\pi ^{}`$ states, which is the most difficult part of the job. We will present the matrix elements of 4-fermion operators as the product of matrix elements of two 2-fermion operators between $`\overline{B}_d`$ and a light meson and vacuum and another light meson. The validity of this factorization is questionable; in particular, in this approach the FSI phases due to the light meson rescattering vanish identically. However, we found the statement in the literature that the corrections to the factorization formulas are small, being proportional to $`\mathrm{\Lambda }/m_b`$ or powers of $`\alpha _S(m_b)`$ <sup>3</sup><sup>3</sup>3Since $`q^2(P_{B_d}P_{\pi ,\rho })^2=O(m_\pi ^2,m_\rho ^2)`$ one can argue that the large distance contributions invalidate the factorization formula.. In any case nowadays factorization is the only way to get expressions for the decay amplitudes from the fundamental Hamiltonian. Since we are interested in $`\overline{B}_d`$ decays to charged mesons and we will factorize 4-fermion operators, let us present Eqs. (A4), (A5) in the following form : $`\widehat{H}`$ $`=`$ $`{\displaystyle \frac{G_F}{\sqrt{2}}}V_{ub}V_{ud}^{}\{a_1\overline{u}\gamma _\alpha (1+\gamma _5)b\overline{d}\gamma _\alpha (1+\gamma _5)u`$ $``$ $`{\displaystyle \frac{V_{tb}V_{td}^{}}{V_{ub}V_{ud}^{}}}[a_4\overline{u}\gamma _\alpha (1+\gamma _5)b\overline{d}\gamma _\alpha (1+\gamma _5)u`$ $``$ $`2a_6\overline{u}(1+\gamma _5)b\overline{d}(1\gamma _5)u]\},`$ where $`a_1=c_1+\frac{1}{3}c_2=1.04`$, $`a_4=c_4+\frac{1}{3}c_3=0.031`$, $`a_6=c_6+\frac{1}{3}c_5=0.042`$ and Fierz identities $`\overline{\psi }\gamma _\alpha (1+\gamma _5)\phi \overline{\chi }\gamma _\alpha (1+\gamma _5)\eta =\overline{\psi }\gamma _\alpha (1+\gamma _5)\eta \overline{\chi }\gamma _\alpha (1+\gamma _5)\phi `$, $`\overline{\psi }\gamma _\alpha (1+\gamma _5)\phi \overline{\chi }\gamma _\alpha (1\gamma _5)\eta =2\overline{\psi }(1\gamma _5)\eta \overline{\chi }(1+\gamma _5)\phi `$ were used. The matrix elements we are interested in were calculated in paper assuming factorization. Up to a common factor which includes constant $`f_\pi `$ and $`B\pi `$ transition formfactor $`f_0(m_\pi ^2)`$ for the amplitude of $`\overline{B}_d\pi ^+\pi ^{}`$ decay it was obtained: $$\frac{M(\overline{B}_d\pi ^+\pi ^{})}{V_{ub}V_{ud}^{}}a_1\frac{V_{tb}V_{td}^{}}{V_{ub}V_{ud}^{}}\left[a_4+\frac{2m_\pi ^2}{(m_u+m_d)(m_bm_u)}a_6\right]e^{i\delta },$$ (A7) where we use the result of and take into account the difference of the rescattering phases of the tree ($`a_1`$) and the penguin ($`a_4`$ and $`a_6`$) amplitudes $`\delta `$. One evident source of this phase is the imaginary part of the penguin diagrams with intermediate $`u`$\- and $`c`$-quarks. Let us demonstrate that only the last one should be taken into account in (A7): $`M(\overline{B}_d`$ $``$ $`\pi ^+\pi ^{})V_{ub}V_{ud}^{}(T+P(m_u))+V_{cb}V_{cd}^{}P(m_c)+V_{tb}V_{td}^{}P(m_t)=`$ (A8) $`=`$ $`V_{ub}V_{ud}^{}[T+P(m_u)P(m_c)]V_{tb}V_{td}^{}[P(m_c)P(m_t)].`$ As we are interested in CP asymmetries we should calculate $`\lambda _{\pi ^+\pi ^{}}=e^{2i\beta }M(\overline{B}_d\pi ^+\pi ^{})/M(B_d\pi ^+\pi ^{})`$: $`\lambda `$ $`=`$ $`e^{2i\beta 2i\gamma }{\displaystyle \frac{1+\frac{P(m_u)P(m_c)}{T}\frac{V_{tb}V_{td}^{}}{V_{ub}V_{ud}^{}}\frac{P(m_c)P(m_t)}{T}}{1+\frac{P(m_u)P(m_c)}{T}\frac{V_{tb}^{}V_{td}}{V_{ub}^{}V_{ud}}\frac{P(m_c)P(m_t)}{T}}}=`$ (A9) $`=`$ $`e^{2i\alpha }\left[1+(e^{i(\alpha \pi )}e^{i(\pi \alpha )}){\displaystyle \frac{\mathrm{sin}\gamma }{\mathrm{sin}\beta }}{\displaystyle \frac{P(m_c)P(m_t)}{T}}\right],`$ where $`\alpha `$, $`\beta `$ and $`\gamma `$ are the angles of the unitarity triangle. Thus the absorptive part of $`P(m_c)`$ contributes to $`\delta `$. Since the penguin operator $`P(m_c)`$ equals the correlator of two vector currents, one immediately picks up its imaginary part from the textbooks on QED. It depends on the gluon momentum transfer and when the square of this momentum transfer is much larger than $`4m_c^2`$, we have: $$\frac{P(m_c)}{T}\mathrm{ln}\frac{M_W^2}{m_b^2}i\pi \left|\frac{P}{T}\right|e^{i\delta },\delta 30^o.$$ (A10) Since $`u`$\- and $`\overline{u}`$-quarks to which gluon decays go to different light mesons, the value of the momentum transfer squared is determined by these mesons wave functions. It varies between $`m_b^2`$ and zero. Thus we see that the mechanism suggested in leads to the small positive value of $`\delta `$: $$\delta \begin{array}{c}<\hfill \\ \hfill \end{array}30^o.$$ (A11) If BSS mechanism determines the value of $`\delta `$, it would confirm the validity of our approach. If, on the contrary, the large distance rescattering of light hadrons changes $`\delta `$ substantially, one should await large corrections to the dispersive part (the ratio ($`P/T`$)) as well. In the present paper we will allow $`\delta `$ to vary between zero and $`2\pi `$, but we use the expressions analogous to (A7) for the decay amplitudes. Let us return to Eq. (A7). With the help of the following equation: $$\frac{V_{td}^{}V_{tb}}{V_{ud}^{}V_{ub}}=e^{i(\pi \alpha )}\frac{\mathrm{sin}\gamma }{\mathrm{sin}\beta }$$ (A12) and using the numerical values $`m_u+m_d=11`$ MeV, $`m_b=4.5`$ GeV we obtain: $`M(\overline{B}_d\pi ^+\pi ^{})`$ $``$ $`V_{ub}V_{ud}^{}\left[1e^{i(\pi \alpha )}{\displaystyle \frac{\mathrm{sin}\gamma }{\mathrm{sin}\beta }}(0.06)e^{i\delta }\right]=`$ (A13) $`=`$ $`V_{ub}V_{ud}^{}[1+0.14e^{i(\pi \alpha +\delta )}],`$ where $`\beta =23^o`$ and $`\gamma =63^o`$ were substituted (we are using the value of $`\gamma `$ from the global CKM fit in order to estimate a small correction to the amplitude) and we put $`a_1`$ equal to one. In Section 4 we analyze the experimental data on CP asymmetries in $`\overline{B}_d(B_d)\pi ^+\pi ^{}`$ decays. Coming to $`\overline{B}_d(B_d)\rho ^+\rho ^{}`$ decays we should calculate the corresponding matrix element of the Hamiltonian presented in (A6). Factorizing 4-quark operators we observe that the term proportional to $`a_6`$ vanishes, since $`<\rho |\overline{d}(1\gamma _5)u|0>=0`$: the (pseudo) scalar current cannot produce a vector meson from vacuum. That is why instead of Eq. (A7) we get (see also ): $`{\displaystyle \frac{M(\overline{B}_d\rho _L^+\rho _L^{})}{V_{ub}V_{ud}^{}}}`$ $``$ $`a_1{\displaystyle \frac{V_{tb}V_{td}^{}}{V_{ub}V_{ud}^{}}}a_4e^{i\delta },`$ $`M(\overline{B}_d\rho _L^+\rho _L^{})`$ $``$ $`V_{ub}V_{ud}^{}[1+0.07e^{i(\pi \alpha +\delta )}].`$ (A14) The production of transversely polarized $`\rho `$-mesons by the vector current is suppressed as $`(m_\rho /m_B)^2`$, and the experimental data confirm the dominance of $`\rho _L`$. If BSS mechanism is responsible for the phase $`\delta `$, then it should be the same as in Eq. (A13). Eq. (A14) is used in Section 3 to extract the value of angle $`\alpha `$. Our last problem is the calculation of the amplitudes $`M(\overline{B}_d\rho ^{}\pi ^\pm )\overline{M}^\pm `$. The amplitude $`\overline{M}^+`$ corresponds to $`\rho ^{}`$ production from vacuum by $`(\overline{d}u)`$ current, so the term proportional to $`a_6`$ does not contribute to it: $`{\displaystyle \frac{\overline{M}^+}{V_{ub}V_{ud}^{}}}`$ $`=`$ $`A\left[a_1{\displaystyle \frac{V_{tb}V_{td}^{}}{V_{ub}V_{ud}^{}}}a_4e^{i\delta _{}}\right];`$ $`\overline{M}^+`$ $`=`$ $`AV_{ub}V_{ud}^{}\left[1+0.07e^{i(\pi \alpha +\delta _{})}\right],`$ (A15) where $`A`$ is the complex number. In the case of the amplitude $`\overline{M}^+`$ it is $`\pi ^{}`$ which is produced from vacuum by $`(\overline{d}u)`$ current, so the term proportional to $`a_6`$ contributes as well: $$\frac{\overline{M}^+}{V_{ub}V_{ud}^{}}=B\left\{a_1\frac{V_{tb}V_{td}^{}}{V_{ub}V_{ud}^{}}\left[a_4\frac{2m_\pi ^2}{(m_b+m_u)(m_u+m_d)}a_6\right]e^{i\delta _+}\right\},$$ (A16) see . Here $`B`$ is the complex number. Unlike the case of $`\overline{B}_d\pi ^+\pi ^{}`$ decay the terms proportional to $`a_4`$ and $`a_6`$ have opposite signs and as a result the expression in square brackets with good accuracy equals zero, leading to: $$\overline{M}^+=BV_{ub}V_{ud}^{},$$ (A17) and the penguin pollution is absent ($`a_1`$ is omitted since it is very close to one). The amplitudes of $`B_d`$ meson decays, $`M^+`$ and $`M^+`$, equal to $`\overline{M}^+`$ and $`\overline{M}^+`$ correspondingly with the complex conjugate CKM matrix elements. We will use formulas (A15) and (A17) in order to determine angle $`\alpha `$ in Section 2 and will omit index “–” from $`\delta _{}`$, since $`\delta _+`$ did not enter Eq. (A17). Thus the penguin pollution is minimal in $`\rho \pi `$ mode, intermediate in $`\rho \rho `$ mode and maximal in $`\pi \pi `$ mode (as it was noted in ).
warning/0506/nucl-ex0506019.html
ar5iv
text
# Scaling properties of azimuthal anisotropy of mesons and baryons at RHIC ## 1 Introduction During the early stages of an ultra-relativistic heavy-ion collision, an extremely high energy density system, possibly consisting of a new phase of nuclear matter, is expected to be formed . The dynamical evolution of this matter is predicted to reflect its properties . Consequently, much effort is currently centered on the study of reaction dynamics at the Relativistic Heavy Ion Collider (RHIC). Azimuthal correlation measurements constitute an important probe for reaction dynamics. They serve as a “barometric sensor” for pressure gradients developed in the collision and hence yield insight into crucial issues of thermalization and the equation of state (EOS) . They provide important constraints for the density of the medium and the effective energy loss of partons which traverse it . They can even provide valuable information on the gluon saturation scale in the nucleus . Azimuthal correlation measurements show significant harmonic strength at mid-rapidity with characteristic dependencies on $`p_T`$ and centrality . The anisotropy of this harmonic pattern is typically characterized by the second order Fourier coefficient, $`v_2=e^{i2(\varphi _1\mathrm{\Phi }_{RP})},`$ (1) where $`\varphi _1`$ represents the azimuthal emission angle of an emitted particle and $`\mathrm{\Phi }_{RP}`$ is the azimuth of the reaction plane. The brackets denote statistical averaging over particles and events. A large amount of available data for Au+Au collisions at $`\sqrt{s_{NN}}`$ = 62.4, 130, and 200 GeV indicate that the magnitude and trends of $`v_2`$ (for $`p_T\stackrel{<}{}2.0`$ GeV/$`c`$) are under-predicted by hadronic cascade models supplemented with string dynamics , but are well reproduced by models which incorporate hydrodynamic flow . This has been interpreted as evidence for the production of a thermalized state of partonic matter . If this is indeed the case, then the fine structure of azimuthal anisotropy (ie. its detailed dependence on centrality, transverse momentum, particle type, higher harmonics, etc) should reflect the scaling “laws” predicted by ideal hydrodynamics. In this work we use detailed differential $`v_2`$ measurements to test for such scaling “laws” and the onset of competing mechanisms. ## 2 Hydrodynamic scaling An important scaling prediction of hydrodynamic theory is exemplified by the exact analytic hydro solutions exploited in the Buda-Lund model . For harmonic flow, the model gives: $$v_{2n}=\frac{I_n(w)}{I_0(w)},n=1,2,..,w=\frac{p_t^2}{4\overline{m}_t}(\frac{1}{T_y}\frac{1}{T_x}),$$ (2) where $`I_{0,n}`$ are modified Bessel-functions, $`\overline{m}_t`$ is an average of the rapidity dependent transverse mass (at mid-rapidity, $`\overline{m_T}=m_T`$), and $`T_x`$ and $`T_y`$ are direction ($`x`$ and $`y`$) dependent slope parameters: $`T_x`$ $`=`$ $`T_0+\overline{m}_t\dot{X}_f^2{\displaystyle \frac{T_0}{T_0+\overline{m}_ta}},`$ (3) $`T_y`$ $`=`$ $`T_0+\overline{m}_t\dot{Y}_f^2{\displaystyle \frac{T_0}{T_0+\overline{m}_ta}}.`$ (4) Here, $`\dot{X}_f`$ and $`\dot{Y}_f`$ gives the transverse expansion rate of the fireball at freeze-out, and $`a=(T_0T_s)/T_s`$ is its transverse temperature inhomogeneity, characterized by the temperature at its center $`T_0`$, and at its surface $`T_s`$. The important prediction of Eq. 2 is that the relatively complicated fine structure of azimuthal anisotropy can be scaled to a single function. An illustration of this fact can be made for particle flavor scaling via substitution of the transverse rapidity , $`y_T=sinh^1(p_T/m)`$ in Eq. 2 to give $$v_2=k_1y_T^2\frac{m}{T_0}\left(1+\frac{k_2}{k_1}\frac{T_0}{m}+\frac{k_3}{k_1}\frac{T_0^2}{m}\mathrm{}\right),$$ (5) where, k1, k2, k3… are largely governed by the expansion rate. Close inspection of the leading term in Eq. 5 indicates that $`v_2`$ for different particle species should scale with $`y_T^{fs}=k_my_T^2m`$. Here $`k_m`$ is a mass dependent factor which is $`1`$ for relatively heavy particles. It is straightforward to show that hydrodynamics also predicts that $`v_2`$ should scale with the spatial eccentricity $`\epsilon =(Y^2X^2)/(Y^2+X^2)`$, of the overlap between the two colliding gold nuclei and the higher harmonic $`v_4\frac{1}{2}v_2^2+k_my_T^4`$. ## 3 Analysis The present analysis is based on $`22`$ M minimum bias Au+Au events obtained with the PHENIX detector at $`\sqrt{s_{NN}}`$ = 200 GeV during the second running period (2002) at RHIC. Charged tracks were detected in the east and west central arms of PHENIX , each of which subtends 90 in azimuth $`\varphi `$, and $`\pm 0.35`$ units of pseudo-rapidity $`\eta `$. Track reconstruction was accomplished at each collision energy via pattern recognition using a drift chamber (DC) followed by two layers of multi-wire proportional chambers with pad readout (PC1 and PC3) located at radii of 2 m, 2.5 m and 5 m respectively . The collision vertex $`z`$ along the beam direction was constrained to be within $`|z|<`$ 30 cm. A confirmation hit within a $`2\sigma `$ matching window was required in PC3 and the electromagnetic calorimeter (EMC PbSc) or the time-of-flight detector (TOF), to eliminate most albedo, conversions, and decays. Particle momenta were measured with a resolution of $`\delta p/p=0.7\%1.0\%p`$ (GeV/$`c`$). Event centralities were obtained via a series of cuts in the space of BBC versus ZDC analog response; they reflect percentile cuts on the total interaction cross section . Estimates for the number of participant nucleons N<sub>part</sub>, were also made for each of these cuts following the Glauber-based model detailed in Ref. . In this analysis the combination of the TOF detector and six sectors of the (EMC PbSc) was used to identify charged particles. Particle time-of-flight was measured using the TOF ( or EMC ) and the collision time defined by beam counters (BBC). A timing resolution of $``$ 120 ps and $``$ 370-400 ps was obtained for the TOF and the EMC (PbSc) respectively. This allowed meson ( $`\pi ^\pm `$, K<sup>±</sup> ) and baryon ( p , $`\overline{p}`$ ) separation up to a $`p_T`$ 3.5 GeV/c with the TOF detector and up to 2.5 GeV/c with the (EMC PbSc). The differential $`v_2(p_T,centrality)`$ measurements were obtained via the reaction plane technique which correlates the azimuthal angles of charged tracks detected in the central arms with the azimuth of an estimated event plane $`\mathrm{\Phi }_2`$, determined via hits in the North and South BBC’s located at $`\eta 33.9`$ . Due the large rapidity gap ($`33.9`$ units) between the particles used for reaction plane determination and the mid-rapidity particles correlated with this plane, one expects that the latter correlations are less influenced by non-flow contributions especially for $`p_T<3.0`$ GeV/$`c`$. The estimated resolution of the combined reaction plane from both BBC’s has an average of 0.33 over centrality with a maximum of about 0.42 for Au+Au collisions at $`\sqrt{s_{NN}}`$= 200 GeV. Thus, the estimated correction factor, which is the inverse of the resolution for the combined reaction plane, ranges from 2.4 to 5.0. ### 3.1 $`v_2`$ results Figures 1 and 2 summarize the centrality and $`p_T`$ dependence of $`v_2`$ for charged mesons( $`\pi ^\pm `$, K<sup>±</sup> ) and baryons ( p + $`\overline{p}`$ ) detected in the EMC+TOF respectively. They give an excellent overview of the the evolution of $`v_2`$ as centrality and $`p_T`$ are varied. The results shown for protons and anti-protons give an especially good view of the evolution away from the well know quadratic dependence of $`v_2(p_T)`$ (which is also observed in very central collisions for these data) as the collisions become more peripheral. Such a dependence could result from changes in the freeze-out temperatures and/or the radial flow velocity as the collisions become more peripheral. ### 3.2 Eccentricity scaling As indicated earlier, hydrodynamics predicts eccentricity scaling of the azimuthal anisotropy. To test for this scaling, one can divide the $`v_2`$ values obtained at a given centrality by the eccentricity $`ϵ`$ obtained from a Glauber-based calculation obtained for the same centrality . Alternatively, one can simply divide the differential anisotropy $`v_2(N_{part},p_T)`$ obtained at a given centrality by the $`p_T`$ integrated value for the same centrality selection $`v_2(N_{part})`$. The underlying idea here is that the integral flow is monotonic and linearly related to the eccentricity over a broad range of centralities . Another advantage of this approach is that the ratio $`v_2(N_{part},p_T)`$/$`v_2(N_{part})`$ gives a scale invariant variable which automatically reduces the systematic errors associated with an eccentricity evaluation, and the reaction plane resolution. Fig. 3 shows the v<sub>2</sub> of charged mesons obtained for different centralities scaled by the value of the integral flow obtained for each of these centralities. The figure shows essentially perfect scaling for mesons as would be expected for a process driven largely by the eccentricity of the overlap region of the two colliding nuclei. The results of a similar scaling test for baryons are shown in Fig. 4. Despite a continuous evolution in the shape of v<sub>2</sub> vs $`p_T`$ with centrality, they indicate a relatively good scaling for the two centrality ranges: 0-20% ( left panel ) and 20-50% ( right panel ). ### 3.3 Flavor scaling Following Eq. 5, the scaling variable $`y_T^{fs}=k_my_T^2m`$, was used to investigate flavor scaling. The results of such a test is summarized in Fig. 5. The left panel of the figure shows a comparison of $`v_2(p_T)`$ for protons, kaons and pions measured with the TOF detector for the centrality selection 5-30%. The characteristic flavor dependence of $`v_2`$ is cleanly evidenced by these data, ie. mass ordering at low momentum and a reversal of the magnitudes of the $`v_2`$ values for baryons and mesons for $`p_T`$ 1.8-2.0 GeV/c. The right panel of Fig. 5 and Fig. 6 show that very good scaling of $`v_2`$ is achieved with $`y_T^{fs}=k_my_T^2m`$ in accordance with the predictions of hydrodynamics. Fig. 6 gives an illustration of y$`{}_{}{}^{fs}{}_{T}{}^{}`$ scaling for combined results which include the neutral kaons and lambda hyperons measured by the STAR collaboration. The latter were obtained in Au+Au collisions at $`\sqrt{s_{NN}}`$ =200 GeV for the same centrality selection . Although the data shown in Fig. 6 indicate rather good scaling for all measured particles over a broad range of y$`{}_{}{}^{fs}{}_{T}{}^{}`$, one can see clear evidence for a break in this scaling for y$`{}_{T}{}^{fs}2`$. Such a break could be signaling a change in mechanism for high $`p_T`$ particles or a break down of ideal hydrodynamic flow. ### 3.4 Higher harmonic scaling As was pointed out earlier, ideal hydrodynamics gives predicts a straightforward scaling relationship between the second harmonic $`v_2`$ and the higher harmonics of azimuthal anisotropy ( $`v_4`$, $`v_6`$, …). It is easy to show that the leading term of the relationship between $`v_2`$ and $`v_4`$ can can be expressed as $$v_4=\frac{1}{2}v_2^2+k_5y_T^4,$$ (6) The measurement of the higher harmonics of azimuthal anisotropy are not available in PHENIX ( ongoing ). However, we can test the predicted scaling relationship between $`v_2`$ and $`v_4`$ via data published by the STAR collaboration . Figure 7 shows the $`(y_T^{fs})^2`$ dependence of $`v_4`$ for charged particles from Au+Au collisions at $`\sqrt{s_{NN}}`$ =200 GeV and $`v_2`$ values scaled according to Eq. 6. The figure indicates rather good agreement between the measured $`v_4`$ values and scaled $`v_2`$ values over a broad range in $`y_T^{fs}`$. ## 4 Summary and Conclusions In summary detailed measurements of the fine structure of elliptic flow in Au+Au collisions at $`\sqrt{s_{NN}}`$ =200 GeV measured by PHENIX collaboration at RHIC are presented. They show eccentricity scaling and $`y_T^{fs}`$ flavor scaling over a broad range of centralities and particle flavors. This observed scaling gives strong support for essentially ideal hydrodynamical flow at RHIC. The observed scaling also supports the picture of a suddenly hadronizing (recombining) fluid of quarks. ## Acknowledgment(s) The author thanks R.A. Lacey, M. Csanád and T. Csörgő for stimulating discussions.
warning/0506/math-ph0506013.html
ar5iv
text
# Deformed 𝐶_𝜆-Extended Heisenberg Algebra in Non-commutative Phase-Space ## Abstract We construct a deformed $`C_\lambda `$-extended Heisenberg algebra in two-dimensional space using non-commuting coordinates which close an algebra depends on statistical parameter characterizing exotic particles. The obtained symmetry is nothing but an exotic particles algebra interpolating between bosonic and deformed fermionic algebras. ## 1 Introduction The recent upsurge of interest in the physics in non-commutative spaces has been spurred due to its very clear and cogent appearance in the context of string theories, D-branes and M-theories. In quantum field theory, this is motivated by studies of the low energy effective theory of D-brane with a non-zero NS-NS B field background . On other hand, the study at the level of quantum mechanics in non-commutative spaces is also meaningful for clarifying some possible phenomenological consequences in solvable models. In this latter context, the present paper is basically devoted to deal with exotic particles living in two-dimensional space, such that a consistent ansatz of commutation relations of phase-space variables should simultaneously include space-space non-commutativity and momentum-momentum non-commutativity. Then, we obtain a new type of commutation relations at the deformed level defining the exotic particles algebra and we show that the obtained symmetry is nothing but a deformed $`C_\lambda `$-extended Heisenberg algebra which is kind of deformed oscillator algebra. As known in the literature, the subject of oscillator algebras has been a topic of intensive research activities during the past decade and much attention has been paid to this field in connection with the important role investigated in many physical systems, such as the description of the fractional statistics . Among these various deformations and extensions, we mention the generalized deformed oscillator algebras (GDOA’s) and the $`G`$-extended oscillator algebras where $`G`$ is some finite group, e.g in the case of Calogero-model $`C_\lambda =𝒵_\lambda `$ is the cyclic group of order $`\lambda `$. The goal of the present letter is mainly to obtain an algebra describing the planar system (exotic particles), which is different from that given by Lerda and Sciuto in the ref. . The study is based on the non-commutative geometry defined by a fundamental algebra depending on the statistical parameter $`\nu `$ which characterizes exotic particles. In section 2, we give a short review on these latter particles. Then we recall some facts concerning the $`C_\lambda `$-extended oscillator algebras which were introduced as a generalization of Calogero-Vasiliev algebras in section 3. In section 4, we consider a non-commuting spatial and momentum coordinates satisfying an algebra depending on the statistical parameters to define an annihilation and creation operators which generate an exotic particles algebra. Owing to the latter algebra. We also show that the obtained symmetry is interpolating between bosonic and deformed fermionic one for arbitrary operator $`\xi `$ which is introduced to define the generators. Another important result is that the algebra describing the planar system is a defomed $`C_\lambda `$-extended Heisenberg algebra. ## 2 Anyons These particles are known as quasi-particles or excitations in two-dimensional space obey intermediate statistics that interpolate between bosonic and fermionic statistics because of its multiply connected configuration space . Among the theories describing anyons there is Chern-Simons theory and generalized Maxwell theory . In the first theory, the standard way to obtain anyons is to add to the action $`S`$ a topological, or Hopf, term $$SS+i\nu S_{top}.$$ (1) The simplest case, to describe free anyons, corresponds to $`S`$ being free charged bosons or fermions with U(1) current $`J^\mu `$ and charges $`Q=\pm 1`$. Here $`S_{top}`$ is the Chern-Simons term where the $`U(1)`$ gauge field $`A_\mu `$ is manufactured from the U(1) matter current $`J^\mu =ϵ^{\mu \nu \alpha }_\nu A_\alpha `$. The statistical parameter $`\nu `$ is normalized so that the spin of the particle is $`\frac{\nu }{2\pi }`$. The second theory is a novel way to describe anyons without a Chern-Simons term. Thus, a generalized connection was considered in (2+1)-dimensions denoted $`A_\mu ^\theta `$, $`\mu =0,1,2`$. The gauge theory is defined by the following Lagrangian $$L_\theta =\frac{1}{2}F_{\mu \nu }F^{\mu \nu }+J^\mu A_\mu ^\theta $$ (2) with $`A_\mu ^\theta A_\mu +\frac{\theta }{2}ϵ_{\mu \nu \rho }F^{\nu \rho }`$ is the generalized connection and $`\theta `$ is real parameter in Minkowski space. The Lagrangian $`L_\theta `$ desribes Maxwell theory that couples to the current via the generalized connection rather than the usual connection. In this model, the gauge field is dynamic and the potential has confining nature which make the theory different from the first one. On other hand, at the level of quantum mechanics in non-commutative spaces, the noncommutative geometry and anyons are related as was shown in the references . The noncommutativity comes from the presence of the magnetic field. In this sense it is valuable to study a field theory both in the presence of a magnetic field and in the coordinate space. In this context, the present work is devoted to find out the symmetry describing anyons basing on noncommutative geometry. We found it very interesting by considering the algebra closed by the coordinates depends on statistical parameter as we will see in the next section. The obtained algebra is interpolating between bosonic and deformed fermionic algebras depending on statistical parameter. After this short review on planar system, we give in what follows its associated symmetry having two extremes: bosonic and deformed fermionic symmetries. ## 3 Planar System Symmetry First, we briefly recall the non-commutative geometry. Its most simple example consists of the geometric space described by non-commutative hermitian operator coordinates $`x_i`$, and by considering the non-commutative momentum operators $`p_i=i_{x_i}`$ $`(_{x_i}`$ the corresponding derivative of $`x_i`$). These operators satisfy the following algebra $$\begin{array}{cc}[x_i,x_j]=i\theta ϵ_{ij},[p_i,p_j]=i\theta ^1ϵ_{ij},[p_i,x_j]=i\delta _{ij}& \\ & \\ [p_i,t]=0=[x_i,t],[p_i,_t]=0=[x_i,_t],& \end{array}$$ (3) with $`t`$ the physical time and $`_t`$ its corresponding derivative. By considering two-dimensional harmonic oscillator which can be decomposed into one-dimensional oscillators. So, it is known that the algebra (2) allows to define, for each dimension, the representation of annihilation and creation operators as follows $$\begin{array}{cc}a_i\hfill & =\sqrt{\frac{\mu \omega }{2}}(x_i+\frac{i}{\mu \omega }p_i)\hfill \\ a_i^{}\hfill & =\sqrt{\frac{\mu \omega }{2}}(x_i\frac{i}{\mu \omega }p_i).\hfill \end{array}$$ (4) with $`\mu `$ is the mass and $`\omega `$ the frequency. These operators satisfy $$[a_i,a_i^{}]=1,$$ defining the Heisenberg algebra. In the simultanuously non-commutative space-space and non-commutative momentum-momentum, the bosonic statistics should be maintained; i.e, the operators $`a_i^{}`$ and $`a_j^{}`$ are commuting for $`ij`$. Thus, the deformation parameter $`\theta `$ is required to satisfy the condition $$\theta =(\frac{1}{\mu \omega })^2\theta ^1.$$ To find out an algebra describing the planar system we start by introducing the non-commutative geometry depending on the statistical parameter $`\nu 𝐑`$. We study two cases with the main difference is based on the deformation of the commutative relation $$[p_i,x_j]=i\delta _{ij}.$$ (5) ### 3.1 Case 1 In this case, we deform the commutation (5). Thus, the fundamental algebra is defined by the spatial coordinates $`x_i`$ and the momentum $`p_i`$ satisfying Proposition 1 $$\begin{array}{cc}[x_i,x_j]_\chi =i\theta ϵ_{ij},[p_i,p_j]_\chi =i\theta (\mu \omega )^2ϵ_{ij},[p_i,x_j]_\chi =i\eta \delta _{ij}& \\ & \\ [p_i,t]=0=[x_i,t],[p_i,_t]=0=[x_i,_t],& \end{array}$$ (6) with the second deformation parameter $`\chi `$ is given by Definition 1 $$\chi =e^{\pm i\nu \pi },$$ (7) where $`\pm `$ sign indicates the two rotation directions on two-dimensional space. $`\theta `$ and $`\eta `$ are a non-commutative parameters depending on statistical parameter $`\nu `$ as we will see later and the notation $`[x,y]_q=xyqyx`$. Thus, owing to the third equality in (6) we get $$[x_i,p_j]_\chi =(\chi ^1\chi )p_jx_i+i\chi ^1\eta \delta _{ij}.$$ Then, we introduce an operator $`\xi _i`$ acting on the momentum direction in the phase-space. We assume that $`\xi _i`$ satisfies the following commutation relation Proposition 2 $$[\xi _i,x_j]=0i,j.$$ (8) In this case, we define the annihilation and the creation operators by Definition 2 $$\begin{array}{cc}b_i^{}\hfill & =\sqrt{\frac{\mu \omega }{2}}(x_i+\frac{i}{\mu \omega }\xi _ip_i)\hfill \\ & \\ b_i^+\hfill & =\sqrt{\frac{\mu \omega }{2}}(x_i\frac{i}{\mu \omega }\xi _i^1p_i),\hfill \end{array}$$ (9) with $`\xi _i`$ is defined in terms of statistical parameter $`\nu `$ and an operaor $`K_i`$ which could be a function of the number operator $`N`$ Definition 3 $$\xi _i=e^{i\nu \pi K_i},$$ (10) According to (6-7) the non-commutative geometry leads to a deformed Heisenberg algebra satisfied by the operators (9) and defined by the following commutation relations $$\begin{array}{ccc}[b_i^{},b_j^+]_\chi \hfill & =\frac{1}{2}\eta (\xi _i\chi ^1+\xi _j^1)\delta _{ij}+i\frac{\mu \omega }{2}\theta (I+\xi _i\xi _j^1)ϵ_{ij}\frac{i\xi _j^1}{2}B_{ij},\hfill & \\ & & \\ [b_i^+,b_j^+]_\chi \hfill & =\frac{1}{2}\eta (\xi _j^1\xi _i^1\chi ^1)\delta _{ij}+i\frac{\mu \omega }{2}\theta (I\xi _i^1\xi _j^1)ϵ_{ij}\frac{i\xi _j^1}{2}B_{ij},\hfill & \\ & & \\ [b_i^{},b_j^{}]_\chi \hfill & =\frac{1}{2}\eta (\xi _i\chi ^1\xi _j)\delta _{ij}+i\frac{\mu \omega }{2}\theta (I\xi _i\xi _j)ϵ_{ij}+\frac{i\xi _j}{2}B_{ij},\hfill & \end{array}$$ (11) with $`I`$ is the identity and $`B_{ij}=(\chi ^1\chi )p_jx_i`$. To be consistent with the hermiticty of coordinates we suggest that $`\theta `$ and $`\eta `$ are operators satisfying $$\theta ^{}=\chi ^1\theta ,\eta ^{}=\chi ^1\eta $$ and we assume that there is some function $`f:𝐑𝐑`$ such that $$\begin{array}{ccc}\hfill f^{}(\nu )=f(\nu ),& f(0)1& \end{array}$$ and $$\underset{\nu 1}{lim}\frac{1+\chi }{f(\nu )}=2.$$ Then, we give the following expressions Definition 4 $$\begin{array}{cc}\theta =\nu (1+\chi )I\hfill & \\ & \\ \eta =\frac{1}{2}\frac{1+\chi }{f(\nu )}\mathrm{cos}\nu 2\pi ,\hfill & \end{array}$$ (12) which are compatible with the commutative phase-space at the extremes concerning the statistical parameter $`\nu `$; $`\theta =0`$ and $`\eta =1`$ for $`\nu =0,1`$ respectively. We remark that the algebra (11) is a deformed version of Heisenberg algerbra satisfied by the operators given in (3). This new algebra describes the anyonic system for arbitrary statistical parameter $`\nu `$. Another important point is that the obtained deformed Heisenberg algerbra (11) is interpolating between two extremes depending on the statistical parameter $`\nu `$. We know that, in three or more dimensions, $`\nu `$ takes the values 0 or 1 and in two dimensions $`\nu `$ is arbitrary real number. The latter case has already discussed above and characterizing exotic particles. In the case of three or more dimensions if $`\nu =0`$ we get $`\chi =1`$, $`\xi _i=\xi _i^1=I`$, $`\theta =0`$, $`\eta =1`$ and $`B_{ij}=0`$. Thus, the commutation relations of the algebra (11) becomes $$\begin{array}{ccc}[b_i^{},b_j^+]=\delta _{ij},\hfill & [b_i^+,b_j^+]=0,& \hfill [b_i^{},b_j^{}]=0.\end{array}$$ (13) These relations define the bosonic algebra and this is one extreme. The second extreme could be gotten if $`\nu =1`$, then we have $`\chi =1`$, $`\theta =0`$, $`\eta =1`$ and $`B_{ij}=0`$. We find the following commutation relations $$\begin{array}{ccc}\{b_i^{},b_j^+\}\hfill & =\frac{1}{2}(e^{i\pi K_j}e^{i\pi K_i})\delta _{ij},\hfill & \\ & & \\ \{b_i^+,b_j^+\}\hfill & =\frac{1}{2}(e^{i\pi K_j}+e^{i\pi K_i})\delta _{ij},\hfill & \\ & & \\ \{b_i^{},b_j^{}\}\hfill & =\frac{1}{2}(e^{i\pi K_i}+e^{i\pi K_j})\delta _{ij}.\hfill & \end{array}$$ (14) which close a deformed fermionic algebra for arbitrary operator $`K_i`$ (10) as a second extreme for the algebra (11). Again if $`K_i`$ is a hermitian operaor then $`\xi _i`$ is unitary and $`b_i^+`$ is a complex conjugate of $`b_i^{}`$, then the algebra (11) will not have the fermionic algebra as extreme when $`\nu =1`$ but its two extremes are bosonic algebra and deformed fermionic algebra. ### 3.2 Case 2 In this subsection, let us see what will happen if we don’t deform the combined commutator of momentum and spatial coordinates. If we keep the commutation relation (5) plus the other relations of the equation (6) the noncommutative geometry is now defined by the following fundamental algebra Prposition 3 $$\begin{array}{cc}[x_i,x_j]_\chi =i\theta ϵ_{ij},[p_i,p_j]_\chi =i\theta (\mu \omega )^2ϵ_{ij},[p_i,x_j]_=i\delta _{ij}& \\ & \\ [p_i,t]=0=[x_i,t],[p_i,_t]=0=[x_i,_t].& \end{array}$$ (15) By straightforward calculations we obtain $$\begin{array}{cc}[x_i,p_j]_\chi =i\delta _{ij}+C_{ji},\hfill & \hfill [p_i,x_j]_\chi =i\delta _{ij}+D_{ji},\end{array}$$ (16) where $`C_{ji}=(1\chi )p_jx_i`$ and $`D_{ji}=(1\chi )x_jp_i`$. Consequently the exotic particles algebra becomes $$\begin{array}{ccc}[b_i^{},b_j^+]_\chi \hfill & =\frac{1}{2}(\xi _i+\xi _j^1)\delta _{ij}+i\frac{\mu \omega }{2}\theta (I+\xi _i\xi _j^1)ϵ_{ij}\frac{i}{2}(\xi _j^1C_{ji}\xi _iD_{ji}),\hfill & \\ & & \\ [b_i^+,b_j^+]_\chi \hfill & =\frac{1}{2}(\xi _j^1\xi _i^1)\delta _{ij}+i\frac{\mu \omega }{2}\theta (I\xi _i^1\xi _j^1)ϵ_{ij}\frac{i}{2}(\xi _j^1C_{ji}+\xi _i^1D_{ji}),\hfill & \\ & & \\ [b_i^{},b_j^{}]_\chi \hfill & =\frac{1}{2}(\xi _i\xi _j)\delta _{ij}+i\frac{\mu \omega }{2}\theta (I\xi _i\xi _j)ϵ_{ij}+\frac{i}{2}(\xi _jC_{ji}+\xi _iD_{ji}),\hfill & \end{array}$$ (17) with $`\theta `$ is defined by (12). Again, if $`\nu =0`$ we get $`\chi =1`$, $`\theta =0`$ and $`C_{ji}=0=D_{ji}`$ and we refind the bosonic algebra (13) as an extreme of the symmetry (17) describing quasi-particles system. Then the case $`\nu =1`$ leads to $`\chi =1`$, $`\theta =0`$ and $`C_{ji}0D_{ji}`$ and we find a deformed fermionic algebra which is different from (14) and defined by $$\begin{array}{ccc}\{b_i^{},b_j^+\}=\frac{1}{2}(e^{i\pi K_i}+e^{i\pi K_j})\delta _{ij}\frac{i}{2}(e^{i\pi K_j}C_{ji}e^{i\pi K_i}D_{ji}),& & \\ & & \\ \{b_i^+,b_j^+\}=\frac{1}{2}(e^{i\pi K_j}e^{i\pi K_i})\delta _{ij}\frac{i}{2}(e^{i\pi K_j}C_{ji}+e^{i\pi K_i}D_{ji}),& & \\ & & \\ \{b_i^{},b_j^{}\}=\frac{1}{2}(e^{i\pi K_i}e^{i\pi K_j})\delta _{ij}+\frac{i}{2}(e^{i\pi K_j}C_{ji}+e^{i\pi K_i}D_{ji}).& & \end{array}$$ (18) as a second extreme of (17). The main result we get from this investigation is that exotic particles algebra goes to bosonic algebra if $`\nu 0`$. This means that our system is originally gotten by exciting a bosonic system in two-dimensional space. Also, we get a deformed fermionic algebra as a second extreme when the statistical parameter $`\nu `$ equals to 1. Thus, we remark that the system described by the above algebras (11) or (17) doesn’t have any thing to do with fermions originally but it could be related to something else as deformed fermions which are known in the literature as quionic particles or $`k_i`$-fermions, $`k_i`$ integer number introduced as deformation parameter, and these kinds of particles are known as non physical particles. ## 4 Deformed Oscillator Algebras In this section we show that the extended Heisenberg algebra could be a symmetry of planar system at defomed level. First we start by a short review on $`C_\lambda `$-extended oscillator algebra and then we give the defomed form of this symmetry which describes exotic particles in two-dimensional space. ### 4.1 Extended Heisenberg Algebra We review in brief the $`C_\lambda `$-extended oscillator algebras. As known in the literature, a generalization of the Calogero-Vasiliev algebras, the $`C_\lambda `$-extended oscillator algebras (also called generalized deformed oscillator algebras (GDOA’s)), denoted $`A^\lambda `$, $`\lambda =2,3,\mathrm{},`$ are defined by $$\begin{array}{cccc}[N,a^{}]=a^{},\hfill & \hfill [a,a^{}]=I+\underset{\mu =0}{\overset{\lambda 1}{}}\alpha _\mu P_\mu ,& & \\ & & & \\ [N,P_\mu ]=0,\hfill & \hfill a^{}P_\mu =P_{\mu +1}a^{},& & \end{array}$$ (19) together with their hermitian conjugates, and $$\begin{array}{cccc}P_\mu =\frac{1}{\lambda }\underset{\nu =0}{\overset{\lambda 1}{}}e^{\frac{2\pi i\nu (N\mu )}{\lambda }},\hfill & \hfill \underset{\mu =0}{\overset{\lambda 1}{}}P_\mu =1,& P_\mu P_\nu =\delta _{\mu ,\nu }P_\nu \hfill & \\ & & & \\ \underset{\mu =0}{\overset{\lambda 1}{}}\alpha _\mu =0,\hfill & \hfill \underset{\nu =0}{\overset{\mu 1}{}}\alpha _\nu >1,& \mu =1,\mathrm{},\lambda 1.\hfill & \end{array},$$ (20) where $`\alpha _\mu 𝐑`$, $`N`$ is the number operator and $`P_\mu `$ are the projection operators on subspaces $`F_\mu =\{|k\lambda \mu |k=0,1,2,\mathrm{}\}`$ of the Fock space $`F`$ which is portioning into $`\lambda `$ subspaces. The operators $`a`$ and $`a^{}`$ are defined by $$\begin{array}{cccc}a^{}a=F(N),\hfill & \hfill aa^{}=F(N+1),& & \end{array}$$ (21) where $`F(N)=N+\underset{\mu =0}{\overset{\lambda 1}{}}\beta _\mu P_\mu `$, $`\beta _\mu =\underset{\nu =0}{\overset{\mu 1}{}}\alpha _\nu `$, which is a fundamental concept of deformed oscillators. Let’s denote the basis states of subspaces $`F_\mu `$ by $`|n=|k\lambda +\mu (a^{})^n|0`$ where $`a|0=0`$, $`|0`$ is the vacuum state. The operators $`a`$, $`a^{}`$ and $`N`$ act on $`F_\mu `$ as follows $$\begin{array}{ccc}\hfill N|n=n|n,& a^{}|n=\sqrt{F(N+1)}|n+1,& a|n=\sqrt{F(N)}|n1.\hfill \end{array}$$ (22) According to these relations, $`a`$ and $`a^{}`$ are the annihilation and the creation operators respectively. Particularly, if $`\lambda =2`$, we have two projection operators $`P_0=\frac{1}{2}(I+(1)^N)`$ and $`P_1=\frac{1}{2}(I(1)^N)`$ on the even and odd subspaces of the Fock space $`F`$, and the relations of (1) are restricted to $$\begin{array}{ccc}\hfill [N,a^{}]=a^{},& [a,a^{}]=I+\kappa K,& \{K,a^{}\}=0,\hfill \end{array}$$ (23) with their hermitian conjugates, where $`K=(1)^N`$ is the Klein operator and $`\kappa `$ is a real parameter. These relations define the so-called Calogero-Vasiliev algebra. The $`C_\lambda `$-extended oscillator algebras are seeing as deformation of $`G`$-extended oscillator algebras, where $`G`$ is some finite group, appeared in connection with $`n`$-particle integrable models. In the former case, $`G`$ is the symmetric group $`S_n`$. So, for two particles $`S_2`$, can be realized in terms of $`K`$ and $`S_2`$-extended oscillator algebra becomes a generalized deformed oscillator algebra (GDOA) also known as the Calogero-Vasiliev or modified oscillator algebra. In the $`C_\lambda `$-extended oscillator algebras, $`GC_\lambda `$ is the cyclic group of order $`\lambda `$, $`C_\lambda =\{1,K,\mathrm{},K^{\lambda 1}\}`$. So, these algebras have a rich structure since they depend upon $`\lambda `$ independent real parameters, $`\alpha _0,\alpha _1,\mathrm{},\alpha _{\lambda 1},`$. ### 4.2 Deformed $`C_\lambda `$-Extended Heisenberg Algebra Other version of anyonic algebra can be obtained by treating the special case of statistical parameter $`\nu [0,1]`$. By using, the Taylor expansion to the operator $`\xi _i`$, the first commutation relation in the algebra (17) can be rewritten in this form $$[b_i^{},b_j^+]_\chi =(I+\mathrm{}_{ij}^{[\nu ]})\delta _{ij}+A_{ij}^{[\nu ]}ϵ_{ij}+Q_{ij}^{[\nu ]},$$ (24) where $$\mathrm{}_{ij}^{[\nu ]}=\underset{\mathrm{}=1}{\overset{\frac{n+1}{2}}{}}\kappa _{\nu ,\mathrm{}}\frac{K_i^{2\mathrm{}1}K_j^{2\mathrm{}1}}{2}+\underset{k=1}{\overset{\frac{m}{2}}{}}\kappa _{\nu ,k}\frac{K_i^{2k}+K_j^{2k}}{2},$$ (25) $$Q_{ij}^{[\nu ]}=\frac{i}{2}\underset{p=0}{\overset{\lambda 1}{}}\frac{(i\nu \pi )^p}{p!}((K_j)^pC_{ji}K_i^pD_{ji})$$ and $$A_{ij}^{[\nu ]}=\frac{i\theta \mu w}{2}\left(I+\underset{\alpha =0}{\overset{\lambda 1}{}}\frac{(i\nu \pi )^\alpha }{\alpha !}(K_iK_j)^\alpha \right),$$ (26) with $`m𝐍`$ is even and $`n𝐍`$ odd such that $`n,m\lambda 1`$ with $`\lambda 𝐍`$ by imposing $$K_i^\lambda =I.$$ The coeffecients $`\kappa _{\nu ,\mathrm{}}`$ and $`\kappa _{\nu ,k}`$ are given in terms of statistical parameter as follows $$\kappa _{\nu ,\mathrm{}}=\frac{(i\nu \pi )^{2\mathrm{}1}}{(2\mathrm{}1)!},\kappa _{\nu ,k}=\frac{(i\nu \pi )^{2k}}{(2k)!}.$$ Then, the last two commutation relations of (17) become $$\begin{array}{cc}[b_i^+,b_j^+]_\chi \hfill & =\underset{\alpha =1}{\overset{\lambda 1}{}}\frac{(i\nu \pi )^\alpha }{\alpha !}\frac{K_j^\alpha K_i^\alpha }{2}\delta _{ij}i\frac{\mu \omega }{2}\theta \underset{\alpha =1}{\overset{\lambda 1}{}}\frac{(i\nu \pi )^\alpha }{\alpha !}(K_i+K_j)^\alpha ϵ_{ij}\hfill \\ & \\ & \frac{i}{2}\underset{p=0}{\overset{\lambda 1}{}}\frac{(i\nu \pi )^p}{p!}(K_j^pC_{ji}K_i^pD_{ji}),\hfill \\ & \\ [b_i^{},b_j^{}]_\chi \hfill & =\underset{\alpha =1}{\overset{\lambda 1}{}}\frac{(i\nu \pi )^\alpha }{\alpha !}\frac{K_i^\alpha K_j^\alpha }{2}\delta _{ij}i\frac{\mu \omega }{2}\theta \underset{\alpha =1}{\overset{\lambda 1}{}}\frac{(i\nu \pi )^\alpha }{\alpha !}(K_i+K_j)^\alpha ϵ_{ij}\hfill \\ & \\ & +\frac{i}{2}\underset{p=0}{\overset{\lambda 1}{}}\frac{(i\nu \pi )^p}{p!}(K_j^pC_{ji}K_i^pD_{ji}).\hfill \end{array}$$ (27) Now if we pose the following Proposition 4 $$K_i=e^{i\frac{2\pi }{\lambda }N_i}$$ with $`N_i`$ is a number operator defined in terms of $`b_i^+`$ and $`b_i^{}`$ as $$b_i^+b_i^{}=f(N),$$ (28) with $`f`$ is some function such that the following relations are satisfied $$\begin{array}{cc}[N_i,b_j^{}]=\delta _{ij}b_i,[N_i,b_j^+]=\delta _{ij}b_i^{},\hfill & \\ & \\ K_ib_j=\delta _{ij}e^{i\frac{2\pi }{\lambda }}b_jK_i,K_ib_j^{}=\delta _{ij}e^{i\frac{2\pi }{\lambda }}b_j^{}K_i.\hfill & \end{array}$$ Thus, It is easy to see that the obtained relations (24) and (27) define a physicswise realization of deformed $`C_\lambda `$-extended Heisenberg algebra on two-dimensional non-commutative space describing exotic particles, where $`C_\lambda `$ is a cyclic group $$C_\lambda =\{I,K_i,K_i^2,K_i^3,\mathrm{},K_i^{\lambda 1}\},\lambda 𝐍.$$ Again once, it is clear that the obtained algebra goes to bosonic symmetry if $`\nu `$ goes to 0. ### 4.3 Fock Representation It is convenient to construct a Fock representation for the algebra (15) underlying the non-commutative geometry by way of the operators (9) obeying (24,27). The Fock space is introduced by the set $$F_i=\{n;n=0,1\}$$ (29) with the states $`n`$ are defined as $$n=\frac{1}{f(\sqrt{n!})}(b_i^+)^n0,n=0,1$$ (30) they are the quantum mechanical states inherent to the non-commutativity (15), with $`0`$ the vacuum state and $`f`$ is some funtion definning the nubmer operator $`N_i`$ as given above in (28). The ”exotic” annihilation and creation operators act on Fock space as $$\begin{array}{cc}b_i^+n=f(\sqrt{n+1})n+1,\hfill & \\ & \\ b_i^{}n=f(\sqrt{n})n1.\hfill & \end{array}$$ (31) Then, Owing to (9,31), the non-commuting spatial and momentum coordinates are acting on Fock space as $$\begin{array}{cc}x_in=\frac{1}{\sqrt{2\mu w}}\left[f(\sqrt{n})n1+f(\sqrt{n+1})n+1\right],\hfill & \\ & \\ p_in=\frac{i\sqrt{2\mu w}}{e^{i\frac{2\pi }{\lambda }n}e^{i\frac{2\pi }{\lambda }n}}\left[f(\sqrt{n})n1+f(\sqrt{n+1})n+1\right],\hfill & \end{array}$$ (32) ## 5 Conclusion We want to mention in conclusion that one of the important results we got studing the noncommutative geometry of two-dimensional space is the crucial role of statistical parameter $`\nu `$. In this investigation, we deformed the fundamental algebra describing the noncommutative geometry to find out the symmetry describes exotic particles living in two-dimensional space. The study leaded to an algebra interpolating between bosonic and deformed fermionic algebras. This means that our system is originally gotten by exciting a bosonic system in two-dimensional space and since the second extreme is a deformed fermionic algebra, the exotic particles system doesn’t have anything to do with fermions originally when the statistical parameter $`\nu `$ goes to 1 but it could be related to something else as deformed fermions which are known in the literature as quionic particles or $`k_i`$-fermions, $`k_i`$ integer number introduced as deformation parameter, and these kinds of particles are not physical particles. Then we looked for other face of exotic particles algebra characterized by small statistical parameters. We obtained a deformed $`C_\lambda `$-Extended Heisenberg Algebra describing quasi-particles. This result is a realization, in physicswise, of a deformed version of $`C_\lambda `$-Extended Heisenberg Algebra in two-dimensional non-commutative space. Acknowledgements: The author would like to thank the Abdus Salam ICTP for the hospitality during the visit in which a part of this work was done.
warning/0506/hep-th0506079.html
ar5iv
text
# A note on the static potential for a D=3 Born-Infeld theory coupled to a new generalized connection ## I Introduction In recent years, the nonlinear Born-Infeld electrodynamics have generated a lot of interest due to their appearance in $`D`$-branes physics. It is a well known fact that, for instance, the low energy dynamics of $`D`$-branes have been described by a nonlinear Born- Infeld type action Tseytlin ; Gibbons . It is worth recalling here that Born and Infeld Born suggested to modify Maxwell electrodynamics to get rid of infinities of the theory. In addition to the string interest, the Born-Infeld theory has also attracted attention from different viewpoints. For example, in connection to noncommutative field theories Gomis , also in magnetic monopoles studies Kim , and possible experimental determination of the parameter that measures the nonlinearity of the theory Denisov . On the other hand, we further recall that systems in $`(2+1)`$ dimensions have been extensively discussed in the last few years Deser ; Dunne ; Khare . This is primarily due to the possibility of realizing fractional statistics, where the physical excitations obeying it are called anyons, which continuously interpolate between bosons and fermions. In this respect, the three-dimensional Chern-Simons gauge theory is the key example, so that Wilczek’s charge-flux composite model of anyon can be implemented Wilczek . Interestingly, we call attention to the fact that recently a novel way to describe anyons has been discussed Itzhaki . In this case, the crucial ingredient for this is the introduction of a generalized connection which permits to realize fractional statistics, such that the Chern-Simons term needs not be introduced. Indeed, by using a Lagrangian which describes Maxwell theory coupled to the current via the generalized connection leads to fractional statistics by the same mechanism, as in the case of the Maxwell-Chern-Simons theory, of attaching a magnetic flux to the electrons Itzhaki ; Gaete1 . We also recall that in recent times a great of deal of attention has been devoted to the concept of duality by its unifying role in physics. As is well known, duality refers to an equivalence between two or more quantum field theories whose corresponding classical theories are different. Consequently, one obtains more information about a theory than is possible by considering a single description. The archetype example on this equivalence is the well known duality between the self-dual and Maxwell-Chern-Simons theories in $`(2+1)`$ dimensions Townsend ; Jackiw , which is obtained via the master Lagrangian approach to show the dynamical equivalence between both theories. Other interesting approaches have been used in different theories, such as the nonlinear $`BF`$ models and higher-order derivative models, to establish the duality mapping Clovis . It is, however, desirable to have some additional check on duality from a somewhat different perspective. We are primarily concerned with the physical content associated to duality. In fact, one can verify the existence of duality between two apparently different theories by computing physical quantities in both theories. If the two answers agree the theories should be dual. This approach is the main focus of this note. With these ideas in mind, by computing the interaction potential between trial external charges, we have showed the equivalence between a $`(2+1)`$-dimensional topologically massive Born-Infeld theory Tripathy ; Harikumar and the Maxwell-Chern-Simons theory with a Thirring interaction term among fermions, in the short distance regime Gaete2 . As already mentioned, the aim of the present paper is to further explore this point. With this in view, we propose to extend the previous analysis to a $`(2+1)`$-dimensional Born-Infeld theory coupled to a generalized connection, which permits fractional statistics. Again, as in the case of the topologically massive Born-Infeld theory, we find that the static potential for this new theory, too, agrees with that of the Maxwell-Chern-Simons theory with a Thirring interaction term among fermions. One is thus lead to the interesting conclusion that the lowest-order modification of the static potential due to the presence of the Born-Infeld term, simulates the effect of including the self-interaction term in the Maxwell-Chern-Simons theory. This is our main result. In this way we establish a new equivalence between both theories, in the hope that this will be helpful to gain a better understanding of effective theories in lower dimensions. As already stated, our objective is to compute explicitly the interaction energy between static pointlike sources for a $`(2+1)`$-dimensional gauge theory containing both the Born- Infeld and the new generalized connection terms. The starting point is the Lagrangian: $$=\beta ^2\left[1\sqrt{1+\frac{1}{2\beta ^2}F_{\mu \nu }F^{\mu \nu }}\right]+\frac{\theta }{2}\epsilon _{\mu \nu \rho }J^\mu F^{\nu \rho }A_\mu J^\mu ,$$ (1) where $`J^\mu `$ is the external current and $`\theta `$ is a parameter with dimension $`M^1`$. The parameter $`\beta `$ measures the nonlinearity of the theory and in the limit $`\beta \mathrm{}`$ the Lagrangian (1) reduces to the Maxwell plus a generalized connection term. In order to handle the square root in the Lagrangian (1), we introduce an auxiliary field $`v`$, such that its equation of motion gives back the original theoryTseytlin . This allows us to rewrite the Lagrangian (1) in the form $$=\beta ^2\left[1\frac{1}{2}v\left(1+\frac{1}{2\beta ^2}F_{\mu \nu }F^{\mu \nu }\right)\frac{1}{2v}\right]+\frac{\theta }{2}\epsilon _{\mu \nu \rho }J^\mu F^{\nu \rho }A_\mu J^\mu .$$ (2) Once this is done, the canonical quantization of this theory from the Hamiltonian point of view is straightforward and follows closely that of Refs. Gaete1 ; Gaete2 . The canonical momenta read $`\mathrm{\Pi }^\mu =vF^{\mu 0}+\theta \epsilon ^{0\mu \nu }J_\nu `$, which result in the usual primary constraint $`\mathrm{\Pi }^0=0`$ and $`p\frac{}{\dot{v}}=0`$. The canonical Hamiltonian following from the above Lagrangian $$\begin{array}{c}\hfill H_C=d^2x\left\{\frac{1}{2v}\left(\mathrm{\Pi }^i\mathrm{\Pi }_i2\theta \epsilon ^{ij}\mathrm{\Pi }_iJ_j+2\theta ^2J_jJ^j\beta ^2\right)+\frac{v}{2}\left(\frac{1}{2}F_{ij}F^{ij}+\beta ^2\right)\right\}+\\ \hfill +d^2x\left\{A_0\left(_i\mathrm{\Pi }^iJ^0\right)\frac{\theta }{2}\epsilon _{ik}F^{ik}J^0+A_kJ^k\beta ^2\right\}.\end{array}$$ (3) Requiring the primary constraint $`\mathrm{\Pi }^0=0`$ to be stationary, we obtain the secondary constraint $`\mathrm{\Gamma }_1(x)_i\mathrm{\Pi }^iJ^0=0`$. The consistency condition for the $`p`$ constraint yields no further constraints and just determines the field $`v`$, $$v=\sqrt{\frac{1\frac{1}{\beta ^2}\left(\mathrm{\Pi }^i\mathrm{\Pi }_i2\theta \epsilon ^{ij}\mathrm{\Pi }_iJ_j+2\theta ^2J^jJ_j\right)}{1+\frac{1}{2\beta ^2}F^{ij}F_{ij}}},$$ (4) which will be used to eliminate $`v`$. The extended Hamiltonian that generates translations in time then reads $`H=H_C+𝑑x\left(c_0(x)\mathrm{\Pi }_0(x)+c_1(x)\mathrm{\Gamma }_1(x)\right)`$, where $`c_0(x)`$ and $`c_1(x)`$ are the Lagrange multipliers. Since $`\mathrm{\Pi }_0=0`$ for all time and $`\dot{A}_0\left(x\right)=[A_0\left(x\right),H]=c_0\left(x\right)`$, which is completely arbitrary, we discard $`A_0\left(x\right)`$ and $`\mathrm{\Pi }_0\left(x\right)`$ because they add nothing to the description of the system. As a result, the Hamiltonian becomes $$H=d^2x\left\{\beta ^2\sqrt{\left(1+\frac{B^2}{\beta ^2}\right)\left(1+\frac{𝐃^2}{\beta ^2}\right)}\beta ^2\theta BJ^0+A_kJ^kc^{}\left(x\right)\left(_i\mathrm{\Pi }^iJ^0\right)\right\},$$ (5) where $`c^{}\left(x\right)=c_1\left(x\right)A_0\left(x\right)`$, $`B^2=\frac{1}{2}F_{ij}F^{ij}`$ and $`𝐃^2=\left(\mathrm{\Pi }_i\mathrm{\Pi }^i+2\theta \epsilon ^{ik}J_i\mathrm{\Pi }_k+2\theta ^2J_iJ^i\right)`$. Since there is one first class constraint $`\mathrm{\Gamma }_1(x)`$ (Gauss’ law), we choose one gauge fixing condition that will make the full set of constraints becomes second class. We choose the gauge fixing condition to correspond to $$\mathrm{\Omega }_2\left(x\right)\underset{C_{\xi x}}{}𝑑z^\nu A_\nu \left(z\right)=_0^1𝑑\lambda x^iA_i\left(\lambda x\right)=0,$$ (6) where $`\lambda `$ $`\left(0\lambda 1\right)`$ is the parameter describing the spacelike straight path between the reference points $`\xi ^k`$ and $`x^k`$ , on a fixed time slice. For simplicity we have assumed the reference point $`\xi ^k=0`$. The choice (6) leads to the Poincaré gauge Gaete2 . With this we obtain the only nontrivial Dirac bracket $$\{A_i(x),\pi ^j(y)\}^{}=g_i^j\delta ^{(2)}\left(xy\right)_i^x_0^1𝑑\lambda x^j\delta ^{(2)}\left(\lambda xy\right).$$ (7) Since we are interested in estimating the lowest-order correction to the interaction energy, we will retain only the leading quadratic term in the expression (5). Thus the Hamiltonian may be written as $$H=d^2x\left\{\frac{1}{2}\left(𝐃^2+B^2\right)\left[1\frac{1}{4\beta ^2}\left(𝐃^2+B^2\right)\right]+\frac{1}{2\beta ^2}B^2𝐃^2\theta BJ^0+A_kJ^k\right\}.$$ (8) Notice that $`𝐃`$ refers to the field $`D^ivE^i=\mathrm{\Pi }^i\theta \epsilon ^{ij}J_j`$. With the help of the Dirac bracket (7), we can now write the Dirac brackets in terms of $`B=\epsilon _{ij}^iA^j`$ and $`D^i=\mathrm{\Pi }^i\theta \epsilon ^{ij}J_j`$ fields, that is, $$\{D_i\left(x\right),D_j\left(y\right)\}^{}=0,$$ (9) $$\{B\left(x\right),B\left(y\right)\}^{}=0,$$ (10) $$\{D_i\left(x\right),B\left(y\right)\}^{}=\epsilon _{ij}_x^j\delta ^{(2)}\left(xy\right).$$ (11) It must now be observed that, unlike the Maxwell-Chern-Simons-Born- Infeld theory Gaete2 , in the present model , the bracket (9) is commutative. It gives rise to the following equations of motion for $`D_i`$ and $`B`$ fields: $$\dot{D}_i\left(x\right)=d^2y\left(1\frac{1}{2\beta ^2}\left(B^2𝐃^2\right)B\theta J^0\right)\{D_i\left(x\right),B\left(y\right)\}^{}+J_i\left(x\right)+d^2yJ^k\left(y\right)_k_0^1𝑑\lambda y_i\delta ^{\left(2\right)}\left(\lambda xy\right),$$ (12) $$\dot{B}\left(x\right)=d^2y\left(1\frac{1}{2\beta ^2}\left(𝐃^2B^2\right)\right)\{B\left(x\right),\frac{1}{2}𝐃^2\left(y\right)\}^{}.$$ (13) In the same way, we write the Gauss law as: $$_iD_L^i=J^0,$$ (14) where $`D_L^i`$ refers to the longitudinal part of $`D^i`$. For $`J^i=0`$, the static electromagnetic fields assume the form $$vB(x)\theta J^0=0,$$ (15) $$\epsilon _{ij}_iD_j\left(x\right)=0.$$ (16) Note that Gauss’ law for the present theory reads $`_iD^i=J^0`$, in other words, $$D^i=^i\left(\frac{J^0}{^2}\right),$$ (17) where $`^2`$ is the two-dimensional Laplacian. It is worthwhile recalling at this point that the field $`v`$ at leading order in $`\beta `$ reads $$v=1+\frac{1}{2\beta ^2}\left(𝐃^2B^2\right).$$ (18) Then using $`D^i=vE^i`$, and for $`J^0(t,𝐱)=e\delta ^{(2)}\left(𝐱\right)`$, the electric field at leading order in $`\beta `$ takes the form $$E^i=\left\{1\frac{1}{2\beta ^2}\left(\frac{e}{2\pi }\frac{x_i}{|𝐱|^2}\right)^2\right\}^i\left(\frac{J^0}{^2}\right),$$ (19) where, as in the case of the topologically massive Born-Infeld theory, we have dropped a divergent factor in expression (19) which does not contribute in the limit $`\beta \theta `$. Now we are prepared to compute the potential energy for a pair of static pointlike opposite charges at $`𝐲`$ and $`𝐲^{}`$ along the lines of Refs. Gaete1 ; Gaete2 . In such a case, we start by considering $$Ve\left(𝒜_0\left(𝐲\right)𝒜_0\left(𝐲^{}\right)\right),$$ (20) where the physical scalar potential $`𝒜_0`$ is expressed in terms of the electric field $$𝒜_0(t,𝐱)=_0^1𝑑\lambda x^iE_i(t,\lambda 𝐱).$$ (21) The following remark deserves to be mentioned: Eq.(21) follows from the vector gauge- invariant field Gaete3 : $$𝒜_\mu \left(x\right)A_\mu \left(x\right)+_\mu \left(_\xi ^x𝑑z^\mu A_\mu \left(z\right)\right),$$ (22) where, as in Eq.(6), the line integral appearing in the above expression is along a spacelike path from the point $`\xi `$ to $`x`$, on a fixed time slice. At the same time, it should be noted that the gauge-invariant variables (22) commute with the sole first class constraint (Gauss’ law), supporting the fact that these fields are physical variables Dirac . Using Eq.(19), we rewrite Eq.(21) as $$𝒜_0(t,𝐱)=_0^1𝑑\lambda \left\{1\frac{1}{2}\left(\frac{e}{2\pi \beta }\right)^2\left(\frac{\lambda x_i}{|\lambda 𝐱|^2}\right)^2\right\}x^i_i^{\lambda 𝐱}\left(\frac{J^0\left(\lambda 𝐱\right)}{_{\lambda 𝐱}^2}\right)$$ (23) where $`J^0`$ is the external current. The static current describing two opposite charges $`e`$ and $`e`$ located at $`𝐲`$ and $`𝐲`$ is then given by $`J^0(t,𝐱)=e\left\{\delta ^{\left(2\right)}\left(𝐱𝐲\right)\delta ^{\left(2\right)}\left(𝐱𝐲^{}\right)\right\}`$. Substituting this back into Eq. (23) we get accordingly $$V=\frac{e^2}{\pi }\mathrm{ln}\left(\mu |𝐲𝐲^{}|\right)\frac{e^4}{16\pi ^3\beta ^2}\frac{1}{|𝐲𝐲^{}|^2},$$ (24) where $`\mu `$ is a cutoff with mass dimension introduced to regularize the potential. As mentioned earlier, this is exactly the result obtained for $`QED_3`$ with a Thirring interaction term among fermions, the so-called generalized Maxwell-Chern-Simons gauge theory Gaete3 ; Ghosh , in the short distance regime. This fact has been noted previously for the topologically massive Born-Infeld theory Gaete2 . Incidentally, it is of interest to notice that the result based on the generalized Maxwell-Chern-Simons theory has been derived from the bosonized version of a $`U(1)`$ gauged massive Thirring model Ghosh . This allows us to interpret the model proposed here as a effective theory which contains quantum effects at the classical level. One thus obtains a similarity between the tree level mechanism exploited here and the bosonization tool used in Ref. Ghosh . This observation and the result (24) are new. We briefly summarize the results so far obtained. Our present model, though very simple, has been found to produce results which strongly simulate the general characteristics of the $`QED_3`$ with a Thirring interaction among fermions, exactly as it happens with the topologically massive Born-Infeld theory. It is quite appealing that there are a class of models which can predict the interaction energy (24). Finally, the methodology presented here provides a physically-based alternative to the usual Wilson loop approach. ## II ACKNOWLEDGMENTS Work supported in part by Fondecyt (Chile) grant 1050546.
warning/0506/math-ph0506004.html
ar5iv
text
# Plane rotations and Hamilton-Dirac mechanics ## 1 Introduction and outline of the paper The *quantum groups* are conventionally constructed via deformations (e.g ). But it is also interesting to consider other methods, e.g canonical and path integral quantizations. Then one has to construct the Lagrangian and Hamiltonian of a group under consideration. The crucial idea of such an approach is that the Euler-Lagrange and Hamilton canonical equations must be the Lie equations of the Lie (transformation) group. In this paper, the canonical formalism for real plane rotations is developed. It is shown that the one-parametric real plane rotation group $`SO(2)`$ can be seen as a toy model of the *Hamilton-Dirac mechanics* with constraints . The Lagrangian and Hamiltonian are explicitly constructed. The Euler-Lagrange and Hamiltonian equations coincide with the Lie equations. Consistency of the constraints is checked. It is also shown that the constraints satisfy the canonical commutation relations (CCR). ## 2 Lagrangian and Lie equations Let $`SO(2)`$ be the rotation group of the real two-plane $`^2`$ . Rotation of the plane $`^2`$ by an angle $`\alpha `$ is given by the transformation $$\{\begin{array}{cc}x^{}=f(x,y,\alpha )x\mathrm{cos}\alpha y\mathrm{sin}\alpha \hfill & \\ y^{}=g(x,y,\alpha )x\mathrm{sin}\alpha +y\mathrm{cos}\alpha \hfill & \end{array}$$ We consider the rotation angle $`\alpha `$ as a dynamical variable and the functions $`f`$ and $`g`$ as *field variables* for $`SO(2)`$. Denote $$\dot{f}_\alpha f,\dot{g}_\alpha g$$ The *infinitesimal coefficients* of the transformation are $$\{\begin{array}{cc}\xi (x,y)\dot{f}(x,y,0)=y\hfill & \\ \eta (x,y)\dot{g}(x,y,0)=x\hfill & \end{array}$$ and the *Lie equations* read $$\{\begin{array}{cc}\dot{f}=\xi (f,g)=g\hfill & \\ \dot{g}=\eta (f,g)=f\hfill & \end{array}$$ Our first aim is to find such a Lagrangian $`L(f,g,\dot{f},\dot{g})`$ that the Euler-Lagrange equations $$\frac{L}{f}\frac{}{\alpha }\frac{L}{\dot{f}}=0,\frac{L}{g}\frac{}{\alpha }\frac{L}{\dot{g}}=0$$ correspondingly coincide with the Lie equations. ###### Definition 1 (Lagrangian). The Lagrangian $`L`$ for $`SO(2)`$ can be defined by $$L(f,g,\dot{f},\dot{g})\frac{1}{2}(f\dot{g}\dot{f}g)\frac{1}{2}\left(f^2+g^2\right)$$ ###### Theorem 2. The Euler-Lagrange equations of $`SO(2)`$ coincide with its Lie equations. ###### Proof. Calculate $`{\displaystyle \frac{L}{f}}`$ $`={\displaystyle \frac{}{f}}\left[{\displaystyle \frac{1}{2}}(f\dot{g}\dot{f}g){\displaystyle \frac{1}{2}}\left(f^2+g^2\right)\right]={\displaystyle \frac{1}{2}}\dot{g}f`$ $`{\displaystyle \frac{L}{\dot{f}}}`$ $`={\displaystyle \frac{}{\dot{f}}}\left[{\displaystyle \frac{1}{2}}(f\dot{g}\dot{f}g){\displaystyle \frac{1}{2}}\left(f^2+g^2\right)\right]={\displaystyle \frac{1}{2}}g{\displaystyle \frac{}{\alpha }}{\displaystyle \frac{L}{\dot{f}}}={\displaystyle \frac{1}{2}}\dot{g}`$ from which it follows $$\frac{L}{f}\frac{}{\alpha }\frac{L}{\dot{f}}=0\frac{1}{2}\dot{g}f+\frac{1}{2}\dot{g}=0\dot{g}=f$$ Analogously calculate $`{\displaystyle \frac{L}{g}}`$ $`={\displaystyle \frac{}{g}}\left[{\displaystyle \frac{1}{2}}(f\dot{g}\dot{f}g){\displaystyle \frac{1}{2}}\left(f^2+g^2\right)\right]={\displaystyle \frac{1}{2}}\dot{f}g`$ $`{\displaystyle \frac{L}{\dot{g}}}`$ $`={\displaystyle \frac{}{\dot{g}}}\left[{\displaystyle \frac{1}{2}}(f\dot{g}\dot{f}g){\displaystyle \frac{1}{2}}\left(f^2+g^2\right)\right]={\displaystyle \frac{1}{2}}f{\displaystyle \frac{}{\alpha }}{\displaystyle \frac{L}{\dot{g}}}={\displaystyle \frac{1}{2}}\dot{f}`$ from which it follows $$\frac{L}{g}\frac{}{\alpha }\frac{L}{\dot{g}}=0\frac{1}{2}\dot{f}g\frac{1}{2}\dot{f}=0\dot{f}=g\mathit{}$$ ## 3 Physical interpretation The system of Lie equations is equivalent to the following one: $$\ddot{f}+f=0=\ddot{g}+g$$ The Lagrangian of the latter reads $$L(f,g,\dot{f},\dot{g})\frac{1}{2}\left(\dot{f}^2+\dot{g}^2\right)\frac{1}{2}\left(f^2+g^2\right)$$ The quantity $$T\frac{1}{2}\left(\dot{f}^2+\dot{g}^2\right)$$ is the *kinetic energy* of a moving point $`(f,g)^2`$, meanwhile $$lf\dot{g}g\dot{f}$$ is its *kinetic momentum* with respect to origin $`(0,0)^2`$. By using the Lie equations one can easily check that $$\dot{f}^2+\dot{g}^2=f\dot{g}g\dot{f}$$ This relation has a simple explanation in the kinematics of a rigid body . The kinetic energy of a point can be represented via its kinetic momentum as follows: $$\frac{1}{2}\left(\dot{f}^2+\dot{g}^2\right)=T=\frac{l}{2}=\frac{1}{2}(f\dot{g}g\dot{f})$$ This relation explains the equivalence of the Lagrangians. Both Lagrangians give rise to the same extremals. But one must remember that this relation holds only on the extremals, i.e for the given Lie equations. ## 4 Hamiltonian and Hamilton equations Our aim is to develop canonical formalism for the Lie equations. According to canonical formalism, define the *canonical momenta* as $`p`$ $`{\displaystyle \frac{L}{\dot{f}}}={\displaystyle \frac{}{\dot{f}}}\left[{\displaystyle \frac{1}{2}}(f\dot{g}\dot{f}g){\displaystyle \frac{1}{2}}\left(f^2+g^2\right)\right]={\displaystyle \frac{g}{2}}`$ $`s`$ $`{\displaystyle \frac{L}{\dot{g}}}={\displaystyle \frac{}{\dot{g}}}\left[{\displaystyle \frac{1}{2}}(f\dot{g}\dot{f}g){\displaystyle \frac{1}{2}}\left(f^2+g^2\right)\right]=+{\displaystyle \frac{f}{2}}`$ Note that the canonical momenta do not depend on velocities and so we are confronted with a *constrained system* with two *constraints* $$\phi _1(f,g,p,s)p+\frac{g}{2}=0,\phi _2(f,g,p,s)s\frac{f}{2}=0$$ ###### Definition 3 (Hamiltonian). According to Dirac , the *Hamiltonian* $`H`$ for $`SO(2)`$ can be defined by $`H`$ $`\stackrel{H^{}}{\stackrel{}{p\dot{f}+s\dot{g}L}}+\lambda _1\phi _1(f,g,p,s)+\lambda _2\phi _2(f,g,p,s)`$ $`=p\dot{f}+s\dot{g}L+\lambda _1\left(p+{\displaystyle \frac{g}{2}}\right)+\lambda _2\left(s{\displaystyle \frac{f}{2}}\right)`$ where $`\lambda _1`$ and $`\lambda _2`$ are the *Lagrange multipliers*. ###### Proposition 4. The Hamiltonian of $`SO(2)`$ can be presented as $$H=\frac{1}{2}\left(f^2+g^2\right)+\lambda _1\left(p+\frac{g}{2}\right)+\lambda _2\left(s\frac{f}{2}\right)$$ ###### Proof. It is sufficient to calculate $`H^{}`$ $`p\dot{f}+s\dot{g}L`$ $`=p\dot{f}+s\dot{g}{\displaystyle \frac{1}{2}}(f\dot{g}\dot{f}g)+{\displaystyle \frac{1}{2}}\left(f^2+g^2\right)`$ $`=\dot{f}\left(p+{\displaystyle \frac{g}{2}}\right)+\dot{g}\left(s{\displaystyle \frac{f}{2}}\right)+{\displaystyle \frac{1}{2}}\left(f^2+g^2\right)`$ $`={\displaystyle \frac{1}{2}}\left(f^2+g^2\right)\mathit{}`$ ###### Theorem 5 (Hamiltonian equations). If the Lagrange multipliers $$\lambda _1=g,\lambda _2=f$$ then the Hamiltonian equations $$\dot{f}=\frac{H}{p},\dot{g}=\frac{H}{s},\dot{p}=\frac{H}{f},\dot{s}=\frac{H}{g}$$ coincide with the Lie equations of $`SO(2)`$. ###### Proof. Really, first calculate $`\dot{f}`$ $`={\displaystyle \frac{H}{p}}={\displaystyle \frac{}{p}}\left[{\displaystyle \frac{1}{2}}\left(f^2+g^2\right)g\left(p+{\displaystyle \frac{g}{2}}\right)+f\left(s{\displaystyle \frac{f}{2}}\right)\right]=g`$ $`\dot{g}`$ $`={\displaystyle \frac{H}{s}}={\displaystyle \frac{}{s}}\left[{\displaystyle \frac{1}{2}}\left(f^2+g^2\right)g\left(p+{\displaystyle \frac{g}{2}}\right)+f\left(s{\displaystyle \frac{f}{2}}\right)\right]=f`$ Similarly calculate $`\dot{p}`$ $`={\displaystyle \frac{H}{f}}={\displaystyle \frac{}{f}}\left[{\displaystyle \frac{1}{2}}\left(f^2+g^2\right)g\left(p+{\displaystyle \frac{g}{2}}\right)+f\left(s{\displaystyle \frac{f}{2}}\right)\right]`$ $`=fs+f=s`$ $`\dot{s}`$ $`={\displaystyle \frac{H}{g}}={\displaystyle \frac{}{g}}\left[{\displaystyle \frac{1}{2}}\left(f^2+g^2\right)g\left(p+{\displaystyle \frac{g}{2}}\right)+f\left(s{\displaystyle \frac{f}{2}}\right)\right]`$ $`=g+p+g=p`$ Now use here the constraints $`p=g/2`$ and $`s=f/2`$ to obtain $$\{\begin{array}{cc}\dot{p}=s\hfill & \\ \dot{s}=p\hfill & \end{array}\{\begin{array}{cc}\frac{1}{2}\dot{g}=\frac{1}{2}f\hfill & \\ +\frac{1}{2}\dot{f}=\frac{1}{2}g\hfill & \end{array}\{\begin{array}{cc}\dot{g}=f\hfill & \\ \dot{f}=g\hfill & \end{array}\mathit{}$$ ###### Remark 6. One must remember that on the constraints must be applied after the calculations of the partial derivatives of $`H`$. ###### Corollary 7. The Hamiltonian of $`SO(2)`$ can be presented in the form $$H=fsgp$$ Then the Hamilton equations coincide with the Lie equations of $`SO(2)`$. ###### Remark 8. Note that our hamiltonian $`H`$ is the *angular momentum* of the point $`(f,g)^2`$. This is natural, because we consider plane rotations: the angular momentum is the generator of the rotations. Hamiltonian obtained from conventional Lagrangian will be the *total energy* $`E`$ $`{\displaystyle \frac{1}{2}}\left(p^2+s^2\right)+{\displaystyle \frac{1}{2}}\left(\dot{f}^2+\dot{g}^2\right)`$ $`={\displaystyle \frac{1}{2}}(fsgp)+{\displaystyle \frac{1}{2}}\left(\dot{f}^2+\dot{g}^2\right)`$ ## 5 Poisson brackets and constraint algebra ###### Definition 9 (observables and Poisson brackets). Sufficiently smooth functions of the canonical varibles are called *observables*. The *Poisson brackets* of the observables $`F`$ and $`G`$ are defined by $$\{F,G\}\frac{F}{f}\frac{G}{p}\frac{F}{p}\frac{G}{f}+\frac{F}{g}\frac{G}{s}\frac{F}{s}\frac{G}{g}$$ ###### Example 10. In particular, one can easily check that $$\{f,p\}=1=\{g,s\}$$ and all other Poisson brackets between canonical variables vanish. ###### Example 11. In particular, $`\{\phi _1,H^{}\}=\{p+{\displaystyle \frac{g}{2}},H^{}\}={\displaystyle \frac{H^{}}{f}}+{\displaystyle \frac{1}{2}}{\displaystyle \frac{H^{}}{s}}={\displaystyle \frac{1}{2}}{\displaystyle \frac{}{f}}\left(f^2+g^2\right)=f`$ and similarly $`\{\phi _2,H^{}\}=\{s{\displaystyle \frac{f}{2}},H^{}\}={\displaystyle \frac{1}{2}}{\displaystyle \frac{H^{}}{p}}{\displaystyle \frac{H^{}}{g}}={\displaystyle \frac{1}{2}}{\displaystyle \frac{}{g}}\left(f^2+g^2\right)=g`$ ###### Definition 12 (weak equality). The observables $`A`$ and $`B`$ are called *weakly equal*, if $$(AB)|_{\phi _1=0=\phi _2}=0$$ In this case we write $`AB`$. ###### Theorem 13. The Lie equations read $$\dot{f}\frac{H}{p},\dot{g}\frac{H}{s},\dot{p}\frac{H}{f},\dot{s}\frac{H}{g}$$ ###### Theorem 14. The Lie equations of $`SO(2)`$ can be presented in the Poisson-Hamilton form $$\dot{f}\{f,H\},\dot{g}\{g,H\},\dot{p}\{p,H\},\dot{s}\{s,H\}$$ ###### Proof. As an example, check the third equation. We have $$\{p,H\}\frac{p}{f}\frac{H}{p}\frac{p}{p}\frac{H}{f}+\frac{p}{g}\frac{H}{s}\frac{p}{s}\frac{H}{g}=\frac{H}{f}\dot{p}\mathit{}$$ ###### Theorem 15. The equation of motion of an observable $`F`$ reads $$\dot{F}\{F,H\}$$ ###### Proof. By using the Hamilton equations, calculate $`\dot{F}`$ $`={\displaystyle \frac{F}{f}}\dot{f}+{\displaystyle \frac{F}{p}}\dot{p}+{\displaystyle \frac{F}{g}}\dot{g}+{\displaystyle \frac{F}{s}}\dot{s}`$ $`{\displaystyle \frac{F}{f}}{\displaystyle \frac{H}{p}}{\displaystyle \frac{F}{p}}{\displaystyle \frac{H}{f}}+{\displaystyle \frac{F}{g}}{\displaystyle \frac{H}{s}}{\displaystyle \frac{F}{s}}{\displaystyle \frac{H}{g}}`$ $`\{F,H\}\mathit{}`$ ###### Theorem 16 (constraint algebra). The constraints of $`SO(2)`$ satsify the CCR relations $$\{\phi _1,\phi _1\}=0=\{\phi _2,\phi _2\},\{\phi _1,\phi _2\}=1$$ ###### Proof. First two relations are evident. To check the third one, calculate $`4\{\phi _1,\phi _2\}`$ $`=\{2p+g,2sf\}`$ $`{\displaystyle \frac{(2p+g)}{f}}{\displaystyle \frac{(2sf)}{p}}{\displaystyle \frac{(2p+g)}{p}}{\displaystyle \frac{(2sf)}{f}}`$ $`+{\displaystyle \frac{(2p+g)}{g}}{\displaystyle \frac{(2sf)}{s}}{\displaystyle \frac{(2p+g)}{s}}{\displaystyle \frac{(2sf)}{g}}`$ $`=2{\displaystyle \frac{(2sf)}{f}}+{\displaystyle \frac{(2sf)}{s}}`$ $`=2+2=4\mathit{}`$ ## 6 Consistency Now consider the dynamical behaviour of the constraints. Note that $$\phi _1=0=\phi _2\dot{\phi _1}=0=\dot{\phi _2}$$ To be consistent with equations of motion we must prove the ###### Theorem 17 (consistency). The constraints of $`SO(2)`$ satisfy equations $$\{\phi _1,H\}\dot{\phi }_1=0,\{\phi _2,H\}\dot{\phi }_2=0$$ ###### Proof. Really, first calculate $`\{\phi _1,H\}`$ $`\{\phi _1,H^{}+\lambda _1\phi _1+\lambda _2\phi _2\}`$ $`\{\phi _1,H^{}\}+\lambda _1\{\phi _1,\phi _1\}+\lambda _2\{\phi _1,\phi _2\}`$ $`=f+\lambda _10+\lambda _21`$ $`=f+f`$ $`=0`$ $`=\dot{\phi }_1`$ Similarly $`\{\phi _2,H\}`$ $`\{\phi _2,H^{}+\lambda _1\phi _1+\lambda _2\phi _2\}`$ $`\{\phi _2,H^{}\}+\lambda _1\{\phi _2,\phi _1\}+\lambda _2\{\phi _2,\phi _2\}`$ $`=g\lambda _11+\lambda _20`$ $`=g+g`$ $`=0`$ $`=\dot{\phi }_2\mathit{}`$ ###### Concluding remark 18. Once the canonical structure of $`SO(2)`$ established, one can perform the canonical quantization as well. This actually means the quantization of the angular momentum. ## Acknowledgement The paper was in part supported by the Estonian SF Grant 5634.
warning/0506/astro-ph0506246.html
ar5iv
text
# Toward a unified light curve model for multi-wavelength observations of V1974 Cygni (Nova Cygni 1992) ## 1 Introduction It has been widely accepted that classical novae are a thermonuclear runaway event on a mass-accreting white dwarf (WD). Characteristic properties on nova evolution have been understood from its ignition through the end of nuclear burning (e.g., Warner, 1995, for a review). The next step we need is quantitative studies of individual objects. For instance, fitting of multi-wavelength light curves with theoretical models enables us to determine nova parameters. Such a work has been developed in the recurrent novae (e.g., Hachisu & Kato, 2001a, b; Hachisu et al., 2000, 2003), but not yet in the classical novae except a pioneering work on V1668 Cyg (Nova Cygni 1978) by Kato (1994). V1974 Cygni (Nova Cygni 1992) is a best example for such studies because it was extensively observed in all the wavelengths from $`\gamma `$-ray to radio. Among various observational data, three bands of optical, continuum ultraviolet (UV) at 1455 Å , and X-ray are used for our present study. Based on an optically thick wind model of nova outbursts (e.g., Kato & Hachisu, 1994), we try to develop a unified model that yields light curves for each wavelength band. The next section introduces the light curve analysis based on our optically thick wind model. In §3, we describe light curve fittings with X-ray, UV, and optical bands. Discussion follows in §4. ## 2 Modeling of V1974 Cyg ### 2.1 Optically thick wind model After a thermonuclear runaway sets in on a mass-accreting WD, its envelope expands greatly to $`R_{\mathrm{ph}}100R_{\mathrm{}}`$ and settles in a steady-state. The decay phase of nova can be followed by a sequence of steady state solutions (e.g., Kato & Hachisu, 1994). Using the same method and numerical techniques as in Kato & Hachisu (1994), we have calculated theoretical light curves. We solve a set of equations, i.e., the continuity, equation of motion, radiative diffusion, and conservation of energy, from the bottom of the hydrogen-rich envelope through the photosphere, under the condition that the solution goes through a critical point of steady-state winds. The winds are accelerated deep inside the photosphere so that they are called “optically thick winds.” We have used updated OPAL opacities (Iglesias & Rogers, 1996). We simply assume that photons are emitted at the photosphere as a blackbody with the photospheric temperature of $`T_{\mathrm{ph}}`$. Physical properties of these wind solutions have already been published (e.g., Hachisu & Kato, 2001a, b, 2004; Hachisu et al., 1996, 1999a, 1999b, 2000, 2003; Kato, 1983, 1997, 1999). It should be noticed that a large number of meshes, i.e., more than several thousands grids, are adopted for the wind solutions in an expanded stage of $`R_{\mathrm{ph}}100R_{\mathrm{}}`$. Optically thick winds stop after a large part of the envelope is blown in the winds. The envelope settles into a hydrostatic equilibrium where its mass is decreasing in time by nuclear burning. Then we solve equation of static balance instead of equation of motion. When the nuclear burning decays, the WD enters a cooling phase, in which the luminosity is supplied with heat flow from the ash of hydrogen burning. ### 2.2 Multiwavelength light curves In the optically thick wind model, a large part of the envelope is ejected continuously for a relatively long period (e.g., Kato & Hachisu, 1994). After the maximum expansion of the photosphere, its photospheric radius gradually decreases keeping the total luminosity ($`L_{\mathrm{ph}}`$) almost constant. The photospheric temperature ($`T_{\mathrm{ph}}`$) increases in time because of $`L_{\mathrm{ph}}=4\pi R_{\mathrm{ph}}^2\sigma T_{\mathrm{ph}}^4`$. The main emitting wavelength of radiation moves from optical to supersoft X-ray through UV. This causes the decrease in optical luminosity and the increase in UV. Then the UV flux reaches a maximum. Finally the supersoft X-ray flux increases after the UV flux decays. These timescales depend on WD parameters such as the WD mass and chemical composition of the envelope (Kato, 1997). Thus, we can follow the development of optical, UV, and supersoft X-ray light curves by a single modeled sequence of steady wind solutions. ### 2.3 System parameters of optically thick wind model The light curves of our optically thick wind model are parameterized by the WD mass ($`M_{\mathrm{WD}}`$), the chemical composition of the envelope, and the envelope mass ($`\mathrm{\Delta }M_{\mathrm{env},0}`$) at the outburst (day 0). We have searched for the best fit model by changing these parameters, for example, in a step of $`0.05M_{\mathrm{}}`$ for the WD mass, of $`0.01`$ for hydrogen mass content, $`X`$, and of 0.05 for carbon, nitrogen, and oxygen mass content, $`C+N+O`$. It should be noted here that hydrogen content $`X`$ and carbon, nitrogen, and oxygen content $`C+N+O`$ are important because they are main players in the CNO cycle but neon content is not because neon is not involved in the CNO cycle. The metal abundance of $`Z=0.02`$ is adopted, in which carbon, nitrogen, oxygen, and neon are also included with the solar composition ratio. We assume neon mass content of $`Ne=0.05`$ (taken from Vanlandingham et al., 2005), because neon mass content cannot be determined only from our light curve fitting. ## 3 Light curve fitting ### 3.1 Supersoft X-ray and UV 1455 Å fluxes ROSAT observation clearly shows that the supersoft X-ray flux emerged on day $`260`$ after the outburst and then decayed rapidly on day $`600`$ through a plateau phase of $`300`$ days (Krautter et al., 1996). Here, we define JD 2,448,665.0 as the outburst day, 8.67 days before the optical maximum (JD 2,488,673.67). We have calculated many models, some of which are plotted in Figure 1 for the wavelength window of $`0.12.4`$ keV. Here, we have determined three parameters of $`M_{\mathrm{WD}}`$, $`X`$, and $`C+N+O`$ by fitting three epochs with the observation, i.e., (1) when wind stops, (2) when hydrogen-burning ends, and (3) when UV 1455 Å flux reaches its maximum. We searched for the best fit model by eye. Our calculated X-ray fluxes in Figure 1 show that the more massive the WD, the shorter the duration of X-ray flat peak, if the other two parameters are the same. This is because a stronger gravity in more massive WDs results in a smaller ignition mass. As a result, hydrogen is exhausted in a shorter period (see, e.g., Kato, 1997, for X-ray turn-off time). On the other hand, if we increase hydrogen content, we have a longer duration of hydrogen burning. In this way, we choose the parameters that fit observed light curves. The best fit model is $`M_{\mathrm{WD}}=1.05M_{\mathrm{}}`$, $`X=0.46`$, $`C+N+O=0.15`$, $`Ne=0.05`$, and $`\mathrm{\Delta }M_{\mathrm{env},0}1.7\times 10^5M_{\mathrm{}}`$, which is denoted by a thick solid line in Figure 1. Two epochs in the best-fit model are indicated by arrows: (1) when the optically thick wind stops and (2) when the steady hydrogen-burning ends. Thin solid lines in Figure 1 depict 1.0, 1.1, and $`1.2M_{\mathrm{}}`$ WDs with $`X=0.35`$, $`C+N+O=0.30`$, and $`Ne=0.0`$ while a thick solid line does the best-fit model of $`1.05M_{\mathrm{}}`$ WD with $`X=0.46`$, $`C+N+O=0.15`$, and $`Ne=0.05`$. To see the effect of hydrogen content, we have added two other models with the same parameters as the best-fit one except hydrogen content: $`X=0.40`$ (dash-dotted line) and $`X=0.53`$ (dotted line). Figure 1 also shows that soft X-rays emerge on day $`250`$. In our model, soft X-rays appear after the wind stops because the wind absorbs soft X-rays (e.g., Southwell et al., 1996; Hachisu & Kato, 2003a, b, c). The optically thick wind stops on day 245 just the time when the supersoft X-rays emerge. After a plateau phase the X-ray flux quickly decreases as shown in Figure 1 because the hydrogen shell-burning ends on day 558. The photospheric radius ($`R_{\mathrm{ph}}`$), temperature ($`T_{\mathrm{ph}}`$), luminosity ($`L_{\mathrm{ph}}`$), and wind mass loss rate (dashed line) are plotted in Figure 2. Our results are roughly consistent with Balman et al.’s (1998) estimates for the photospheric radii and temperatures. We should place the nova at a distance of 2.2 kpc to fit our calculated X-ray flux with Balman et al.’s fluxes. This distance is longer than that derived from the UV fitting in Figure 3. See discussion. Figure 3 depicts UV fluxes in a band of $`\mathrm{\Delta }\lambda =20`$ Å wide centered at $`\lambda =1455`$ Å, taken from Cassatella et al. (2004). The corresponding UV light curves are calculated for each model in Figure 1. Our best fitted $`1.05M_{\mathrm{}}`$ WD model shows an excellent agreement with the observation if we place the nova at a distance of 1.7 kpc. Here we adopt an absorption law given by Seaton (1979), $`A_\lambda =8.3E(BV)=2.65`$, together with an extinction of $`E(BV)=0.32`$ estimated by Chochol et al. (1997). ### 3.2 Optical fluxes We cannot fit the observed visual light curve by our best fitted model or even by other models with other sets of WD mass and envelope chemical composition. Therefore, we interpret that the optical flux is dominated by free-free emission of the optically thin ejecta that exist outside the photosphere. For the free-free emission of optically thin ejecta, optical flux can be roughly estimated as $$F_\lambda N_eN_i𝑑V_{R_{\mathrm{ph}}}^{\mathrm{}}\frac{\dot{M}_{\mathrm{wind}}^2}{r^4}r^2𝑑r\frac{\dot{M}_{\mathrm{wind}}^2}{R_{\mathrm{ph}}}$$ (1) during the optically thick wind phase, where $`F_\lambda `$ is the flux at the wavelength $`\lambda `$, $`N_e`$ and $`N_i`$ the number densities of electron and ion, $`V`$ the volume of the ejecta, $`\dot{M}_{\mathrm{wind}}`$ the wind massless rate. Here, we use the relation of $`\rho _{\mathrm{wind}}=\dot{M}_{\mathrm{wind}}/4\pi r^2v_{\mathrm{wind}}`$, and $`\rho _{\mathrm{wind}}`$ and $`v_{\mathrm{wind}}`$ are the density and velocity of the wind, respectively. After the wind stops, we obtain $$F_\lambda N_eN_idV\rho ^2V\frac{M_{\mathrm{ej}}^2}{V^2}V(R^3t^3),$$ (2) (e.g., Woodward et al., 1997), where $`\rho `$ is the density, $`M_{\mathrm{ej}}`$ the ejecta mass (in parenthesis, if $`M_{\mathrm{ej}}`$ is constant in time), $`R`$ the radius of the ejecta ($`VR^3`$), and $`t`$ the time after the outburst. Here, we substitute $`\dot{M}_{\mathrm{wind}}`$ and $`R_{\mathrm{ph}}`$ of our best fit model for those in equation(1). We cannot uniquely specify the constant in equations (1) and (2) because radiative transfer is not calculated outside the photosphere. Instead, we choose the constant to fit the light curve on day 43 denoted by A (on the thick solid line) and on day 245 denoted by B (on the dashed line) in Figure 4. These two light curves represent well the early/late parts of the observational data of AAVSO. Woodward et al. (1997) summarized the optical and infrared (IR) observations of V1974 Cyg and concluded that $`0.55\mu `$m $`V`$, $`1.25\mu `$m $`J`$, $`1.6\mu `$m $`H`$, and $`2.3\mu `$m $`K`$ light curves all showed an abrupt transition from a $`t^{1.5}`$ slope to a $`t^3`$ slope at day $`170`$. This $`t^{1.5}`$ slope is very close to the slope of our free-free light curve until day $`100`$. After the wind stops, we have a slope of $`t^3`$ as shown in Figure 4. This transition probably occurs when the optically thick wind stops. Therefore, our model is very consistent with the temporal optical and IR observations. ## 4 Discussion ### 4.1 Hard X-ray component ROSAT observation shows that hard X-ray flux increases on day $`70100`$ and then decays on day $`270300`$. This hard component is suggested to be shock-origin between ejecta (Krautter et al., 1996). Here we present another idea that these hard X-rays are originated from a shock between the optically thick wind and the companion as described below. V1974 Cyg is a binary system with an orbital period of $`P_{\mathrm{orb}}=0.0812585`$ days (e.g., De Young & Schmidt, 1994; Retter et al., 1997). Paresce et al. (1995) and Retter et al. (1997) estimated the companion mass at $`0.21M_{\mathrm{}}`$ from this orbital period. Using these values we obtain the separation, $`a=0.853R_{\mathrm{}}`$, the effective radii of each Roche lobe, $`R_1^{}=0.444R_{\mathrm{}}`$ and $`R_2^{}=0.215R_{\mathrm{}}`$ for the primary (WD) and the secondary component, respectively. Our optically thick wind model predicts that the companion star emerges from the WD photosphere about day 80 (for the photospheric radius, see Fig. 2). Before day $`80`$, the companion resides deep inside the WD photosphere and we do not detect hard X-rays. After the companion emerges from the WD photosphere, the shock front can be directly observed. The optically thick wind stops on day 245 and we expect that the hard X-ray component decays after that. This hard X-ray flux may show orbital modulations if the inclination angle of binary is large enough. However, Chochol et al. (1997) estimated it at $`i39\mathrm{°}`$. For such a small inclination angle, we are able to see main parts of the shock front at any binary phase, because a bow-shock is formed off the surface of the companion (see, e.g., Shima et al., 1986) and basically optically thin to hard X-rays. Therefore, orbital modulation of hard X-ray flux is hardly observed, which is consistent with the observation (Krautter et al., 1996). Balman et al. (1998) estimated the hydrogen column density of the hard X-ray component and concluded that it decreases, by a factor of $`10`$, from $`N_\mathrm{H}10^{22.2}`$ to $`10^{21.3}`$ cm<sup>-2</sup> between day 70 and day 260, and almost constant after that. In our optically thick wind model, the neutral hydrogen column density is given by $$N_\mathrm{H}\frac{X}{m_\mathrm{H}}_{r_\mathrm{s}}^{\mathrm{}}\rho _{\mathrm{wind}}𝑑r\frac{\dot{M}_{\mathrm{wind}}X}{4\pi av_{\mathrm{wind}}m_\mathrm{H}}\dot{M}_{\mathrm{wind}},$$ (3) where $`m_\mathrm{H}`$ is the mass of hydrogen atom, $`r_\mathrm{s}`$ the position of the bow-shock from the WD center, $`a`$ the separation of the binary, and we roughly assume that the bow-shock front is at distance of $`r_\mathrm{s}=(0.50.7)a`$ from the WD center. Our wind mass loss rate decreases from $`10^5`$ to $`10^6M_{\mathrm{}}`$ yr<sup>-1</sup> between day 70 and day 260 (see Fig. 2), which is very consistent with Balman et al.’s results. ### 4.2 WD mass and chemical composition Several groups estimated the WD mass of V1974 Cyg. Retter et al. (1997) gave a mass of $`M_{\mathrm{WD}}=0.751.07M_{\mathrm{}}`$ based on the precessing disk model of superhump phenomenon. A similar range of $`0.751.1M_{\mathrm{}}`$ is also obtained by Paresce et al. (1995) from various empirical relations on novae. Very recently, Sala & Hernanz (2005) found the WD mass to be $`0.9M_{\mathrm{}}`$ for 50% mixing of a solar composition envelope with a O-Ne degenerate core, or $`1.0M_{\mathrm{}}`$ for 25% mixing, by comparing the evolutional speed of post-wind phase of V1974 Cyg with their post-wind phase of static envelope solutions. Their values for the $`1.0M_{\mathrm{}}`$ are roughly consistent with our results. Vanlandingham et al. (2005) criticized Austin et al.’s (1996) results and reanalyzed chemical abundances of the ejecta from optical and UV spectra. They obtained that He$`=1.2\pm 0.2`$, C$`=0.7\pm 0.2`$, N$`=44.9\pm 11`$, O$`=12.8\pm 7`$, and Ne$`=41.5\pm 17`$ by number relative to hydrogen and relative to solar. In our notation, these correspond to $`X=0.55`$, $`Y=0.25`$, $`C+N+O=0.12`$, $`Ne=0.06`$, and $`Z=0.02`$ by mass weight. The hydrogen content is a bit higher but these values are very consistent with our results. ### 4.3 Distance We estimate the distance to V1974 Cyg from the UV 1455 Å light curve fitting. The absorption at $`\lambda =1455`$ Å is calculated to be $`A_\lambda =8.3E(BV)=2.65`$ (e.g., Seaton, 1979), where we adopt the absorption at the visual band, $`A_V=3.1E(BV)=0.99`$ (Chochol et al., 1997). Then we have a distance to the nova of $`d1.7`$ kpc. For the X-ray band, Balman et al. (1998) obtained $`(R_{\mathrm{ph}}/d)^2=(0.220.26)\times 10^{25}`$ on day 518 (corresponding to their day 511). Using $`R_{\mathrm{ph}}=0.0115R_{\mathrm{}}`$ on day 518 of our best fit model, we obtain the distance of $`d=1.61.7`$ kpc, which is consistent with our distance estimation from the UV fitting. On the other hand, our X-ray flux combined with Balman et al.’s (1998) fluxes gives a rather large distance of 2.2 kpc. This difference may come from the different model parameters adopted in their atmosphere models: $`1.2M_{\mathrm{}}`$ WD and $`X=0.54`$, $`Y=0.21`$, $`Z=0.02`$, $`C=0.002`$, $`O=0.103`$, $`N=0.002`$, and $`Ne=0.123`$. Their neon mass is much higher than the observation $`Ne=0.06`$ (Vanlandingham et al., 2005). Therefore we take the distance of $`d=1.7`$ kpc. These distances are all within the range listed in Chochol et al. (1997), $`d=1.33.5`$ kpc with a most probable value of 1.8 kpc, derived mainly from maximum magnitude-rate of decline (MMRD) relations. We thank A. Cassatella for providing us with their machine readable UV 1455 Å data of V1974 Cygni and also AAVSO for the visual data of V1974 Cygni. We are also grateful to the anonymous referee for useful comments to improve the manuscript. This research has been supported in part by the Grant-in-Aid for Scientific Research (16540211, 16540219) of the Japan Society for the Promotion of Science.
warning/0506/math0506605.html
ar5iv
text
# Convergence of the Wick Star Product ## 1 Introduction Deformation quantization usually comes in two flavours: formal and strict deformations: In formal deformation quantization as introduced by , see also for recent reviews, one considers the Poisson algebra of smooth complex-valued functions $`C^{\mathrm{}}(M)`$ on a Poisson manifold as observable algebra in classical mechanics. A formal star product $``$ is a $`[[\lambda ]]`$-bilinear associative product for $`C^{\mathrm{}}(M)[[\lambda ]]`$ such that in zeroth order $`fg`$ is the pointwise product and in first order of $`\lambda `$ the $``$-commutator gives $`\mathrm{i}`$ times the Poisson bracket. Usually one requires the higher orders to be given by bidifferential operators. The algebra $`C^{\mathrm{}}(M)[[\lambda ]]`$ then serves as a model for the quantum mechanical observables corresponding to the classical system described by $`M`$. In particular, the formal parameter $`\lambda `$ corresponds to Planck’s constant $`\mathrm{}`$ and should be replaced by $`\mathrm{}`$ whenever one can establish convergence of the formal series. On the other hand, in strict deformation quantization as introduced in , see also , one works on the framework of $`C^{}`$-algebras where the dependence of the deformed product $`_{\mathrm{}}`$ on the deformation parameter $`\mathrm{}`$ is now required to be continuous. This is made precise using the notion of continuous fields of $`C^{}`$-algebras. While in the first approach one has very strong existence and classification results , the formal character of Planck’s constant is, of course, physically not acceptable. Here the second approach is much more appealing as it directly uses the analytical framework suitable for quantum mechanics. On the other hand, however, a general construction and reasonable classification of strict quantizations seems still to be missing. Many examples like the global symbol calculus on cotangent bundles , Berezin-Toeplitz quantization on Kähler manifolds as well as suggest that the formal star products should be seen as asymptotic expansions for $`\mathrm{}0`$ of their convergent counterparts in strict quantization. On the other hand, many formal star products allow for large subalgebras, where the formal series actually converge, whence in some sense the asymptotics can be used again to recover the strict result, a heuristic statement for which a general theorem unfortunately is still missing. The above examples also suggest that there is a framework in between formal and $`C^{}`$-algebraic, namely one can try to construct deformations of $`C^{\mathrm{}}(M)`$ (or suitable subalgebras of it) in the framework of *Fréchet* or more generally *locally convex algebras*. Early results in this direction have been obtained in , see also . Moreover, a general set-up of smooth deformations has been established and exemplified in , in holomorphic deformations were studied. The example we are going to discuss will provide an entire holomorphic deformation of a Fréchet subalgebra of $`C^\omega (M)`$. More specifically, we consider the most simple phase space $`M=^n`$ with its canonical Poisson structure $`\{z^k,\overline{z}^{\mathrm{}}\}=\frac{2}{\mathrm{i}}\delta ^{k\overline{\mathrm{}}}`$ and the *formal Wick star product* $$f_{\mathrm{Wick}}g=\underset{r=0}{\overset{\mathrm{}}{}}\frac{(2\lambda )^r}{r!}\underset{i_1,\mathrm{},i_r}{}\frac{^rf}{z^{i_1}\mathrm{}z^{i_r}}\frac{^rg}{\overline{z}^{i_1}\mathrm{}\overline{z}^{i_r}},$$ (1.1) where $`z^1,\mathrm{},z^n`$ are the canonical, global, holomorphic coordinates on $`^n`$. Our convergence scheme to construct the ‘convergent’ subalgebra $`𝒜_{\mathrm{}}`$ is then based on the crucial observation that the Wick star product enjoys a very strong positivity property : every $`\delta `$-functional is a positive $`[[\lambda ]]`$-linear functional in the sense of formal power series. After choosing a point $`p^n`$ and $`\mathrm{}>0`$ this will allow us to construct recursively a system of seminorms for which the Wick star product is continuous, thereby defining our algebra $`𝒜_{p,\mathrm{}}`$. Moreover, we shall construct, via the GNS construction corresponding to the positive functional $`\delta _p`$, a faithful -representation of $`𝒜_{p,\mathrm{}}`$ on a dense subspace of the Bargmann-Fock space giving an interpretation of the $`\delta `$-functionals as coherent states with respect to the Heisenberg group $`𝖧_n`$ acting on $`^n`$. We treat this example in quite some detail as we believe that it may serve as a good starting point for geometric generalizations to Wick star products on Kähler manifolds suitable for a bottom-up approach to . Moreover, in a future project we shall discuss the field-theoretic generalization for infinitely many degrees of freedom. The paper is organized as follows: In Section 2 we briefly recall some basic properties of $`_{\mathrm{Wick}}`$, the formal GNS construction for $`\delta `$-functionals and the Bargmann-Fock space. Section 3 is devoted to the construction of the seminorms, depending on $`p^n`$ and $`\mathrm{}>0`$. This gives the space $`𝒜_{p,\mathrm{}}`$ which is shown to be a subspace of $`C^\omega (^n)`$ with a Fréchet topology. In Section 4 we show that $`𝒜_{p,\mathrm{}}`$ is a subalgebra of $`C^\omega ()`$ such that the pointwise product as well as the Poisson bracket are continuous, i.e. $`𝒜_{p,\mathrm{}}`$ becomes a Fréchet-Poisson algebra. Moreover, we show that the formula (1.1) for $`_{\mathrm{Wick}}`$ actually converges on $`𝒜_{p,\mathrm{}}`$ in the Fréchet topology resulting in a continuous product. This way, $`𝒜_{p,\mathrm{}}`$ becomes a holomorphic deformation. In Section 5 we discuss the dependence on the a priori chosen point $`p`$ and on the value $`\mathrm{}>0`$ of Planck’s constant. It turns out that the translation group acts on $`𝒜_{p,\mathrm{}}`$ by inner -automorphisms whence $`𝒜_{p,\mathrm{}}=𝒜_{\mathrm{}}`$ does not depend on the choice of $`p`$. As a side remark we show that the star exponential, see , of linear functions converges in the topology of $`𝒜_{\mathrm{}}`$. Moreover, for all values $`\mathrm{}>0`$ the algebras $`𝒜_{\mathrm{}}`$ are isomorphic in a canonical way. Finally, in Section 6 we show how the GNS construction yields a -representation of $`𝒜_{\mathrm{}}`$ in the sense of in the Bargmann-Fock space. The action of the Heisenberg group by inner -automorphisms gives easily the coherent states. Acknowledgement: We would like to thank Pierre Bieliavsky, Nico Giulini, Simone Gutt, Nikolai Neumaier, Konrad Schmüdgen, Martin Schlichenmaier and Rainer Verch for valuable discussions on this example. Moreover, S. W. thanks the ULB for hospitality while part of this work was being done. ## 2 Preliminary results In this section we shall collect some well-known results on the Wick star product which we shall use in the sequel, see e.g. . On the classical phase space $`^{2n}^n`$ with standard symplectic form $`\omega =\frac{\mathrm{i}}{2}\mathrm{d}z^k\mathrm{d}\overline{z}^k`$ one defines the *formal Wick star product* by $$f_{\mathrm{Wick}}g=\underset{r=0}{\overset{\mathrm{}}{}}\frac{(2\lambda )^r}{r!}\underset{i_1,\mathrm{},i_r}{}\frac{^rf}{z^{i_1}\mathrm{}z^{i_r}}\frac{^rg}{\overline{z}^{i_1}\mathrm{}\overline{z}^{i_r}},$$ (2.1) where $`f,gC^{\mathrm{}}(^n)[[\lambda ]]`$, the formal parameter $`\lambda `$ corresponds to Planck’s constant $`\mathrm{}`$ without any further prefactors and $`z`$, $`\overline{z}`$ denote the usual global holomorphic/anti-holomorphic coordinates on $`^n`$. Then $`_{\mathrm{Wick}}`$ is known to be an associative star product quantizing the canonical Poisson bracket corresponding to $`\omega `$. It has the separation of variable property in the sense of Karabegov and is in fact the name-giving example of a star product of Wick type in the sense of . Moreover, $`_{\mathrm{Wick}}`$ is Hermitian $$\overline{f_{\mathrm{Wick}}g}=\overline{g}_{\mathrm{Wick}}\overline{f},$$ (2.2) where according to our interpretation of $`\lambda `$ the formal parameter is defined to be real $`\overline{\lambda }=\lambda `$. We shall also frequently make use of multiindex notation: Let $`R=(r_1,\mathrm{},r_n)^n`$ be a multiindex, then one defines $`|R|=r_1+\mathrm{}+r_n`$, $`R!=r_1!\mathrm{}r_n!`$ as well as $`z^R=(z^1)^{r_1}\mathrm{}(z^n)^{r_n}`$ etc. Moreover, we define $`RL`$ if $`r_i\mathrm{}_i`$ for all $`i=1,\mathrm{},n`$. The Wick star product can equivalently be written as $$f_{\mathrm{Wick}}g=\underset{R=0}{\overset{\mathrm{}}{}}\frac{(2\lambda )^{|R|}}{R!}\frac{^{|R|}f}{z^R}\frac{^{|R|}g}{\overline{z}^R}.$$ (2.3) The Wick star product enjoys a very strong positivity property which e.g. the Weyl-Moyal star product does not share: If $`\delta _p:C^{\mathrm{}}(^n)[[\lambda ]][[\lambda ]]`$ denotes the evaluation functional at $`p^n`$ then we have $$\delta _p(\overline{f}_{\mathrm{Wick}}f)=\underset{R=0}{\overset{\mathrm{}}{}}\frac{(2\lambda )^{|R|}}{R!}\left|\frac{^{|R|}f}{\overline{z}^R}(p)\right|^20,$$ (2.4) where the positivity is understood in the sense of the canonical ring ordering of $`[[\lambda ]]`$, see for a detailed discussion on the physical relevance of this notion of positivity. This very strong positivity property is not true for general Hermitian star products: instead one has to add ‘quantum corrections’ to a given classically positive functional (here $`\delta _p`$) in order to obtain a positive functional with respect to the star product. Since positive functionals play the role of *states* the above simple observation implies that for the Wick star product any classical state defines a quantum state *without any quantum correction*, see for a detailed discussion on the general situation. In fact, the Wick star product is used in an essential way for proving that for an arbitrary Hermitian star product one can always construct quantum corrections for a classical state, see . Since $`\delta _p`$ is a positive functional for $`_{\mathrm{Wick}}`$ we have a corresponding GNS representation of the Wick star product algebra $`C^{\mathrm{}}(^n)[[\lambda ]]`$. In fact, this example was one of the first examples of GNS constructions in deformation quantization. Since we need the construction in the following, we briefly review the results from . The Gel’fand ideal $`𝒥_p`$ of $`\delta _p`$ is given by $$𝒥_p=\left\{f\right|\delta _p(\overline{f}_{\mathrm{Wick}}f)=0\}=\left\{f\right|\frac{^{|R|}f}{\overline{z}^R}(p)=0\text{for all}R\}.$$ (2.5) The GNS pre Hilbert space $`_p=C^{\mathrm{}}(^n)[[\lambda ]]/𝒥_p`$ can canonically be identified with $$_{\mathrm{BF}}=[[\overline{y}^1,\mathrm{},\overline{y}^n]][[\lambda ]]$$ (2.6) with the $`[[\lambda ]]`$-valued positive definite inner product $$\varphi ,\psi =\underset{R=0}{\overset{\mathrm{}}{}}\frac{(2\lambda )^{|R|}}{R!}\overline{\frac{^{|R|}\varphi }{\overline{y}^R}(0)}\frac{^{|R|}\psi }{\overline{y}^R}(0),$$ (2.7) where the identification of a class $`\psi _f_p`$ is given by its formal anti-holomorphic Taylor expansion, i.e. $$_p\psi _f\underset{R=0}{\overset{\mathrm{}}{}}\frac{1}{R!}\frac{^{|R|}f}{\overline{z}^R}(p)\overline{y}^R,$$ (2.8) where $`\psi _f`$ denotes the equivalence class of the function $`f`$ in $`_p`$. Then the GNS representation on $`_p`$ defined by $`\pi _p(f)\psi _g=\psi _{f_{\mathrm{Wick}}g}`$ is translated into $$\varrho _p(f)=\underset{R,S=0}{\overset{\mathrm{}}{}}\frac{(2\lambda )^{|R|}}{R!S!}\frac{^{|R|+|S|}f}{z^R\overline{z}^S}(p)\overline{y}^S\frac{^{|R|}}{\overline{y}^R},$$ (2.9) via the unitary map (2.8). In particular, for $`p=0`$ we see that this gives the formal analog of the usual *Bargmann-Fock space* and the *Bargmann-Fock representation*: indeed, recall that the Bargmann-Fock space is the Hilbert space $$_{\mathrm{BF}}=\left\{f\overline{𝒪}(^n)\right|\frac{1}{(2\pi \mathrm{})^n}|f(\overline{z})|^2\mathrm{e}^{\frac{\overline{z}z}{2\mathrm{}}}\mathrm{d}z\mathrm{d}\overline{z}<\mathrm{}\}$$ (2.10) of *anti-holomorphic* functions which are square-integrable with respect to the Gaussian measure, see . Then it is well-known that $`_{\mathrm{BF}}`$ is actually a closed subspace of the $`L^2`$-space for this measure and hence a Hilbert space itself. Moreover, the $`L^2`$-inner product can be evaluated by the same formula (2.7) if one replaces $`\lambda `$ by $`\mathrm{}`$, where the series now converges absolutely. A Hilbert basis for $`_{\mathrm{BF}}`$ is given by the monomials $$𝖾_R(\overline{z})=\frac{1}{\sqrt{(2\mathrm{})^{|R|}R!}}\overline{z}^R.$$ (2.11) From (2.7) we see that the Hilbert space $`_{\mathrm{BF}}`$ can be interpreted as the space of those anti-holomorphic functions whose Taylor coefficients at $`0`$ form a sequence in a (weighted) $`\mathrm{}^2`$-space. ## 3 Construction of the Fréchet space $`𝒜_{p,\mathrm{}}`$ The motivation for our convergence scheme is rather simple: we fix $`\mathrm{}>0`$ and we fix a point $`p^n`$. Then we are looking for a subalgebra of $`C^{\mathrm{}}(^n)[[\lambda ]]`$ such that $`\delta _p(f)`$ converges for $`\lambda =\mathrm{}`$. Though this looks rather innocent at the beginning, we obtain a hierarchy of conditions: For $`f,g`$ in our *algebra* we want $`f_{\mathrm{Wick}}g`$ to be in the algebra as well whence $`\delta _p(f_{\mathrm{Wick}}g)`$ has to converge for $`\lambda =\mathrm{}`$ as well. The idea is now to estimate the convergence of $`\delta _p(f_{\mathrm{Wick}}g)|_{\lambda =\mathrm{}}`$ by $`\delta _p(\overline{f}_{\mathrm{Wick}}f)|_{\lambda =\mathrm{}}`$ and $`\delta _p(\overline{g}_{\mathrm{Wick}}g)|_{\lambda =\mathrm{}}`$ using the Cauchy-Schwarz inequality for the *positive* $`\delta `$-functional. An iteration of this procedure will lead to countably many *unary* conditions for a function $`f`$ to belong to our algebra, which we shall interpret as seminorms determining the algebra. This approach can hence be used as a heuristic motivation for the following definition of the seminorms. ###### Definition 3.1 Let $`R,S^n`$ be multiindices, $`m`$ and $`\mathrm{}=0,\mathrm{},2^m1`$. Then we define recursively for $`fC^{\mathrm{}}(^n)`$ $$\begin{array}{cc}\hfill h_{0,0,R,S}^{p,\mathrm{}}\left(f\right)& =(2\mathrm{})^{|R|+|S|}\left(\frac{^{|R|+|S|}\overline{f}}{z^S\overline{z}^R}_{\mathrm{Wick}}\frac{^{|R|+|S|}f}{\overline{z}^Sz^R}\right)(p)|_{\lambda =\mathrm{}}\hfill \\ & =\underset{N=0}{\overset{\mathrm{}}{}}\frac{(2\mathrm{})^{|R|+|S|+|N|}}{N!}\left|\frac{^{|R|+|S|+|N|}f}{z^R\overline{z}^{N+S}}(p)\right|^2\hfill \end{array}$$ (3.1) and $$h_{m,\mathrm{},R,S}^{p,\mathrm{}}\left(f\right)=\{\begin{array}{cc}\underset{N=0}{\overset{\mathrm{}}{}}\frac{1}{N!}\left(\underset{I=0}{\overset{R}{}}\underset{J=0}{\overset{N+S}{}}\left(\genfrac{}{}{0.0pt}{}{R}{I}\right)\left(\genfrac{}{}{0.0pt}{}{N+S}{J}\right)h_{m1,\mathrm{}/2,I,J}^{p,\mathrm{}}\left(f\right)\right)^2\hfill & \mathrm{}\text{even}\hfill \\ \underset{N=0}{\overset{\mathrm{}}{}}\frac{1}{N!}\left(\underset{I=0}{\overset{R}{}}\underset{J=0}{\overset{N+S}{}}\left(\genfrac{}{}{0.0pt}{}{R}{I}\right)\left(\genfrac{}{}{0.0pt}{}{N+S}{J}\right)h_{m1,(\mathrm{}1)/2,J,I}^{p,\mathrm{}}\left(f\right)\right)^2\hfill & \mathrm{}\text{odd.}\hfill \end{array}$$ (3.2) The precise form of the recursive definition will become either clear by following the above heuristic argument in detail or from the proof of Proposition 3.7: the binomial coefficients arise from the Leibniz rule. Thanks to the positivity (2.4) of the $`\delta `$-functional it is clear that $`h_{0,0,R,S}^{p,\mathrm{}}\left(f\right)`$ either converges absolutely or diverges absolutely to $`+\mathrm{}`$, as it is a series consisting of non-negative terms only. By induction, the same is true for all other $`h_{m,\mathrm{},R,S}^{p,\mathrm{}}\left(f\right)`$. Hence we have $$h_{m,\mathrm{},R,S}^{p,\mathrm{}}\left(f\right)[0,+\mathrm{}],$$ (3.3) where ‘convergence’ is always absolute and does not depend on the order of summation. ###### Definition 3.2 For $`fC^{\mathrm{}}(^n)`$ we define $$f_{m,\mathrm{},R,S}^{p,\mathrm{}}=\sqrt[2^{m+1}]{h_{m,\mathrm{},R,S}^{p,\mathrm{}}\left(f\right)}[0,+\mathrm{}].$$ (3.4) Moreover, we define $$\stackrel{~}{𝒜}_{p,\mathrm{}}=\left\{fC^{\mathrm{}}(^n)\right|f_{m,\mathrm{},R,S}^{p,\mathrm{}}<\mathrm{}\text{for all}m,\mathrm{},R,S\}.$$ (3.5) The first step is now to show that $`\stackrel{~}{𝒜}_{p,\mathrm{}}`$ is a vector space and that the $`_{m,\mathrm{},R,S}^{p,\mathrm{}}`$ are seminorms on $`\stackrel{~}{𝒜}_{p,\mathrm{}}`$. This will be a consequence of the following proposition: ###### Proposition 3.3 The maps $`_{m,\mathrm{},R,S}^{p,\mathrm{}}:C^{\mathrm{}}(^n)[0,+\mathrm{}]`$ enjoy the following properties: 1. $`\alpha f_{m,\mathrm{},R,S}^{p,\mathrm{}}=|\alpha |f_{m,\mathrm{},R,S}^{p,\mathrm{}}`$ for $`\alpha `$. 2. $`f+g_{m,\mathrm{},R,S}^{p,\mathrm{}}f_{m,\mathrm{},R,S}^{p,\mathrm{}}+g_{m,\mathrm{},R,S}^{p,\mathrm{}}`$. 3. $`f_{m1,\mathrm{},R,S}^{p,\mathrm{}}f_{m,2\mathrm{},R,S}^{p,\mathrm{}}`$ and $`f_{m1,\mathrm{},R,S}^{p,\mathrm{}}f_{m,2\mathrm{}+1,R,S}^{p,\mathrm{}}`$. 4. $`f_{m,\mathrm{},0,S}^{p,\mathrm{}}\sqrt[2^{m+2}]{S!}f_{m+1,2\mathrm{},0,0}^{p,\mathrm{}}`$. 5. $`f_{m,\mathrm{},R,0}^{p,\mathrm{}}\sqrt[2^{m+2}]{R!}f_{m+1,2\mathrm{}+1,0,0}^{p,\mathrm{}}`$. 6. $`f_{m,\mathrm{},R,S}^{p,\mathrm{}}\sqrt[2^{m+3}]{R!}\sqrt[2^{m+2}]{S!}f_{m+2,4\mathrm{}+1,0,0}^{p,\mathrm{}}`$. ###### Proof. The first part is clear by a simple induction. For the second part the case $`m=0`$ follows directly from Minkowski’s inequality. Then $`m>0`$ is shown inductively by using again Minkowski’s inequality, for both cases of odd and even $`\mathrm{}`$. The remaining inequalities between the $`f_{m,\mathrm{},R,S}^{p,\mathrm{}}`$ for different values of the parameters are simply obtained by omitting all but one term for a specific $`N`$ in the defining summations of (3.2). For example, in the third part one considers $`N=0`$ and $`I=R`$, $`J=S`$ only, while for the fourth and fifth one uses $`N=S=J`$ and $`N=R=I`$. ∎ Thus the maps $`_{m,\mathrm{},R,S}^{p,\mathrm{}}`$, restricted to $`\stackrel{~}{𝒜}_{p,\mathrm{}}`$, give indeed seminorms. Moreover, the labels $`R`$, $`S`$ play only a minor role thanks to the estimates in the last part of the proposition. This motivates the following definitions. For $`fC^{\mathrm{}}(^n)`$ we define $$f_{m,\mathrm{}}^{p,\mathrm{}}=f_{m,\mathrm{},0,0}^{p,\mathrm{}}$$ (3.6) $$f_m^{p,\mathrm{}}=\underset{0\mathrm{}2^m1}{\mathrm{max}}\left\{f_{m,\mathrm{}}^{p,\mathrm{}}\right\}.$$ (3.7) Then we have the following simple corollary: ###### Corollary 3.4 The set $`\stackrel{~}{𝒜}_{p,\mathrm{}}C^{\mathrm{}}(^n)`$ is a subvector space and the $`_{m,\mathrm{},R,S}^{p,\mathrm{}}`$ are seminorms on $`\stackrel{~}{𝒜}_{p,\mathrm{}}`$. Moreover, the seminorms $`_{m,\mathrm{}}^{p,\mathrm{}}`$ as well as the seminorms $`_m^{p,\mathrm{}}`$ determine the same locally convex topology on $`\stackrel{~}{𝒜}_{p,\mathrm{}}`$. In the following, we shall equip $`\stackrel{~}{𝒜}_{p,\mathrm{}}`$ always with this locally convex topology induced by the seminorms $`_{m,\mathrm{},R,S}^{p,\mathrm{}}`$. However, this topology has one unpleasant feature: it is *non-Hausdorff* as a function $`f`$ whose Taylor expansion at $`p`$ vanishes identically has clearly $`f_{m,\mathrm{},R,S}^{p,\mathrm{}}=0`$ for all parameters $`m,\mathrm{},R,S`$. On the other hand, as one sees already from the seminorm $`_1^{p,\mathrm{}}`$ the functions with *vanishing* $`\mathrm{}`$-jet $`𝗃_p^{\mathrm{}}f`$ at $`p`$ are the only functions with this property. Thus we identify them to be zero in order to have a Hausdorff topology: ###### Definition 3.5 We define $$𝒜_{p,\mathrm{}}=\stackrel{~}{𝒜}_{p,\mathrm{}}/\left\{fC^{\mathrm{}}(^n)\right|𝗃_p^{\mathrm{}}f=0\},$$ (3.8) and equip $`𝒜_{p,\mathrm{}}`$ with the induced locally convex topology determined by the seminorms $`_{m,\mathrm{},R,S}^{p,\mathrm{}}`$ (or equivalently, by the seminorms $`_m^{p,\mathrm{}}`$). Clearly, $`𝒜_{p,\mathrm{}}`$ is now a Hausdorff locally convex topological vector space. The following theorem shows that the abstract quotient can be viewed as a certain subspace of the real-analytic functions on $`^n`$: ###### Theorem 3.6 Let $`\mathrm{}>0`$ and $`p^n`$. 1. $`𝒜_{p,\mathrm{}}`$ is a Hausdorff locally convex topological vector space. 2. Every class $`[f]𝒜_{p,\mathrm{}}`$ has a unique real-analytic representative $`fC^\omega (^n)`$. Therefore, we identify $`𝒜_{p,\mathrm{}}`$ with the corresponding subspace of $`C^\omega (^n)`$ from now on. 3. Every function $`f𝒜_{p,\mathrm{}}`$ has a unique extension to a function $`\widehat{f}𝒪\times \overline{𝒪}(^n\times ^n)`$, i.e. holomorphic in the first and anti-holomorphic in the second argument, such that $$f=\mathrm{\Delta }^{}\widehat{f},$$ (3.9) where $`\mathrm{\Delta }:^nz(z,z)^n\times ^n`$ is the diagonal. 4. Any $`fC^\omega (^n)`$ such that there exist constants $`a,b,c>0`$ with $$\left|\frac{^{|R|+|S|}f}{\overline{z}^Rz^S}(p)\right|ca^{|R|}b^{|S|}$$ (3.10) belongs to $`𝒜_{p,\mathrm{}}`$. In particular $`[z,\overline{z}]𝒜_{p,\mathrm{}}`$. ###### Proof. The first part is clear. For the second, we consider $$f_{1,1,0,0}^{p,\mathrm{}}\frac{1}{R!}(f_{0,0,R,0}^{p,\mathrm{}})^2\frac{(2\mathrm{})^{2|R|+2|S|}}{R!(S!)^2}\left|\frac{^{|R|+|S|}f}{z^R\overline{z}^S}(p)\right|^4$$ for all $`R`$, $`S`$. Thus we obtain $$\left|\frac{^{|R|+|S|}f}{z^R\overline{z}^S}(p)\right|f_{1,1,0,0}^{p,\mathrm{}}\frac{\sqrt[4]{R!}\sqrt{S!}}{\sqrt{(2\mathrm{})^{|R|+|S|}}}.$$ (3.11) But this implies that the series $$\widehat{f}(z,\overline{w})=\underset{R,S=0}{\overset{\mathrm{}}{}}\frac{1}{R!S!}\frac{^{|R|+|S|}f}{z^R\overline{z}^S}(p)(zp)^R(\overline{w}\overline{p})^S$$ converges for all $`z,w^n`$. Thus $`\widehat{f}𝒪\times \overline{𝒪}(^n\times ^n)`$ and clearly $`\mathrm{\Delta }^{}\widehat{f}`$ is in the same equivalence class as $`f`$. This shows the second and third part. Now assume $`fC^\omega (^n)`$ satisfies (3.10). Then $$h_{0,0,R,S}^{p,\mathrm{}}\left(f\right)c^2\mathrm{e}^{2\mathrm{}b^2n}(2\mathrm{}a^2)^{|R|}(2\mathrm{}b^2)^{|S|},$$ whence there are constants $`c_{0,0}=c^2\mathrm{e}^{2\mathrm{}b^2n}`$, $`a_{0,0}=2\mathrm{}a^2`$ and $`b_{0,0}=2\mathrm{}b`$ such that $$h_{0,0,R,S}^{p,\mathrm{}}\left(f\right)c_{0,0}a_{0,0}^{|R|}b_{0,0}^{|S|}.$$ We claim that for all $`m`$, $`\mathrm{}`$ there are constants $`a_{m,\mathrm{}}`$, $`b_{m,\mathrm{}}`$ and $`c_{m,\mathrm{}}`$ such that $$h_{m,\mathrm{},R,S}^{p,\mathrm{}}\left(f\right)c_{m,\mathrm{}}a_{m,\mathrm{}}^{|R|}b_{m,\mathrm{}}^{|S|}.$$ Indeed, a recursive argument shows that $$c_{m,\mathrm{}}=c_{m1,\mathrm{}/2}^2\mathrm{e}^{n(1+b_{m1,\mathrm{}/2})^2},a_{m,\mathrm{}}=(1+a_{m1,\mathrm{}/2})^2,b_{m,\mathrm{}}=(1+b_{m1,\mathrm{}/2})^2$$ for even $`\mathrm{}`$ and $$c_{m,\mathrm{}}=c_{m1,(\mathrm{}1)/2}^2\mathrm{e}^{n(1+a_{m1,(\mathrm{}1)/2})^2},a_{m,\mathrm{}}=(1+b_{m1,(\mathrm{}1)/2})^2,b_{m,\mathrm{}}=(1+a_{m1,(\mathrm{}1)/2})^2$$ for odd $`\mathrm{}`$ will do the job. But then all seminorms of $`f`$ are finite. ∎ Since the polynomials are in $`𝒜_{p,\mathrm{}}`$ we shall make intense use of them. The next proposition gives a first hint on the continuity of the Wick product. Here and in the following we shall use the notation $`_{\mathrm{Wick}}^{\mathrm{}}`$ for the Wick product with $`\lambda `$ being replaced by $`\mathrm{}`$. ###### Proposition 3.7 Let $`fC^{\mathrm{}}(^n)`$ and let $`g[z,\overline{z}]`$ be a polynomial. 1. $`f_{\mathrm{Wick}}^{\mathrm{}}g`$ is a finite sum and thus a well-defined smooth function. 2. $`\overline{f}_{\mathrm{Wick}}^{\mathrm{}}g_{m,\mathrm{},R,S}^{p,\mathrm{}}f_{m+1,2^m+\mathrm{},R,S}^{p,\mathrm{}}g_{m+1,\mathrm{},R,S}^{p,\mathrm{}}`$. ###### Proof. The first part is clear from the explicit form of $`_{\mathrm{Wick}}`$. For the second part we first have by the Cauchy-Schwarz inequality $$\begin{array}{cc}\hfill \left|\overline{f}_{\mathrm{Wick}}^{\mathrm{}}g\right|^2(p)& =\left|\underset{N=0}{\overset{\text{finite}}{}}\frac{(2\mathrm{})^{|N|}}{N!}\frac{^{|N|}\overline{f}}{z^N}(p)\frac{^{|N|}g}{\overline{z}^N}(p)\right|^2\hfill \\ & \left(\underset{N=0}{\overset{\mathrm{}}{}}\frac{(2\mathrm{})^{|N|}}{N!}\left|\frac{^{|N|}\overline{f}}{z^N}(p)\right|^2\right)\left(\underset{N=0}{\overset{\mathrm{}}{}}\frac{(2\mathrm{})^{|N|}}{N!}\left|\frac{^{|N|}g}{\overline{z}^N}(p)\right|^2\right)\hfill \\ & =h_{0,0,0,0}^{p,\mathrm{}}\left(f\right)h_{0,0,0,0}^{p,\mathrm{}}\left(g\right).\hfill \end{array}$$ Now partial derivatives are still *derivations* of $`_{\mathrm{Wick}}`$ and hence of $`_{\mathrm{Wick}}^{\mathrm{}}`$ if one of the functions is a polynomial. This allows the following computation $$\begin{array}{cc}& h_{0,0,R,S}^{p,\mathrm{}}\left(\overline{f}_{\mathrm{Wick}}^{\mathrm{}}g\right)\hfill \\ & =\underset{N=0}{\overset{\mathrm{}}{}}\frac{(2\mathrm{})^{|N|+|R|+|S|}}{N!}\left|\frac{^{|N|+|R|+|S|}}{\overline{z}^{N+S}z^R}\left(\overline{f}_{\mathrm{Wick}}^{\mathrm{}}g\right)\right|^2(p)\hfill \\ & =\underset{N=0}{\overset{\mathrm{}}{}}\frac{(2\mathrm{})^{|N|+|R|+|S|}}{N!}\left|\underset{I=0}{\overset{R}{}}\underset{J=0}{\overset{N+S}{}}\left(\genfrac{}{}{0.0pt}{}{R}{I}\right)\left(\genfrac{}{}{0.0pt}{}{N+S}{J}\right)\frac{^{|I|+|J|}\overline{f}}{z^I\overline{z}^J}_{\mathrm{Wick}}^{\mathrm{}}\frac{^{|N+SJ|+|RI|}g}{z^{RI}\overline{z}^{N+SJ}}\right|^2(p)\hfill \\ & \underset{N=0}{\overset{\mathrm{}}{}}\frac{(2\mathrm{})^{|N|+|R|+|S|}}{N!}\left(\underset{I=0}{\overset{R}{}}\underset{J=0}{\overset{N+S}{}}\left(\genfrac{}{}{0.0pt}{}{R}{I}\right)\left(\genfrac{}{}{0.0pt}{}{N+S}{J}\right)\left|\frac{^{|I|+|J|}\overline{f}}{z^I\overline{z}^J}_{\mathrm{Wick}}^{\mathrm{}}\frac{^{|N+SJ|+|RI|}g}{z^{RI}\overline{z}^{N+SJ}}\right|\right)^2(p)\hfill \\ & \stackrel{()}{}\underset{N=0}{\overset{\mathrm{}}{}}\frac{(2\mathrm{})^{|N|+|R|+|S|}}{N!}\left(\underset{I=0}{\overset{R}{}}\underset{J=0}{\overset{N+S}{}}\left(\genfrac{}{}{0.0pt}{}{R}{I}\right)\left(\genfrac{}{}{0.0pt}{}{N+S}{J}\right)\sqrt{h_{0,0,0,0}^{p,\mathrm{}}\left(\frac{^{|I|+|J|}f}{\overline{z}^Iz^J}\right)}\sqrt{h_{0,0,0,0}^{p,\mathrm{}}\left(\frac{^{|N+SJ|+|RI|}g}{z^{RI}\overline{z}^{N+SJ}}\right)}\right)^2\hfill \\ & =\underset{N=0}{\overset{\mathrm{}}{}}\frac{1}{N!}\left(\underset{I=0}{\overset{R}{}}\underset{J=0}{\overset{N+S}{}}\left(\genfrac{}{}{0.0pt}{}{R}{I}\right)\left(\genfrac{}{}{0.0pt}{}{N+S}{J}\right)\sqrt{h_{0,0,J,I}^{p,\mathrm{}}\left(f\right)}\sqrt{h_{0,0,RI,N+SJ}^{p,\mathrm{}}\left(g\right)}\right)^2\hfill \\ & \underset{N=0}{\overset{\mathrm{}}{}}\frac{1}{N!}\left(\underset{I=0}{\overset{R}{}}\underset{J=0}{\overset{N+S}{}}\left(\genfrac{}{}{0.0pt}{}{R}{I}\right)\left(\genfrac{}{}{0.0pt}{}{N+S}{J}\right)h_{0,0,J,I}^{p,\mathrm{}}\left(f\right)\right)\left(\underset{I=0}{\overset{R}{}}\underset{J=0}{\overset{N+S}{}}\left(\genfrac{}{}{0.0pt}{}{R}{I}\right)\left(\genfrac{}{}{0.0pt}{}{N+S}{J}\right)h_{0,0,I,J}^{p,\mathrm{}}\left(g\right)\right)\hfill \\ & \sqrt{\underset{N=0}{\overset{\mathrm{}}{}}\frac{1}{N!}\left(\underset{I=0}{\overset{R}{}}\underset{J=0}{\overset{N+S}{}}\left(\genfrac{}{}{0.0pt}{}{R}{I}\right)\left(\genfrac{}{}{0.0pt}{}{N+S}{J}\right)h_{0,0,J,I}^{p,\mathrm{}}\left(f\right)\right)^2}\sqrt{\underset{N=0}{\overset{\mathrm{}}{}}\frac{1}{N!}\left(\underset{I=0}{\overset{R}{}}\underset{J=0}{\overset{N+S}{}}\left(\genfrac{}{}{0.0pt}{}{R}{I}\right)\left(\genfrac{}{}{0.0pt}{}{N+S}{J}\right)h_{0,0,I,J}^{p,\mathrm{}}\left(g\right)\right)^2}\hfill \\ & =\sqrt{h_{1,1,R,S}^{p,\mathrm{}}\left(f\right)}\sqrt{h_{1,0,R,S}^{p,\mathrm{}}\left(g\right)},\hfill \end{array}$$ which proves the second part for the case $`m=0`$. Note the necessity that one function (we have chosen $`g`$) is polynomial since otherwise the ‘function’ $`\overline{f}_{\mathrm{Wick}}^{\mathrm{}}g`$ is a priori not defined as a smooth function. The general case is now obtained by a straightforward induction on $`m`$ only using the Cauchy-Schwarz inequality. ∎ ###### Corollary 3.8 The pointwise complex conjugation is a continuous map $`𝒜_{p,\mathrm{}}𝒜_{p,\mathrm{}}`$. We have $$\overline{f}_{m,\mathrm{},R,S}^{p,\mathrm{}}1_{m+1,\mathrm{},R,S}^{p,\mathrm{}}f_{m+1,2^m+\mathrm{},R,S}^{p,\mathrm{}}.$$ (3.11) Note however that the seminorms themselves are not invariant under complex conjugation $`f\overline{f}`$, though the complex conjugation is continuous. We also note that the second part of the proposition already shows some nice continuity properties of the Wick star product, at least if one function is a polynomial. Note however, that the above argument will not extend to arbitrary $`f,g`$ whence we shall need another route. ###### Theorem 3.9 The polynomials $`[z,\overline{z}]𝒜_{p,\mathrm{}}`$ are dense. More specifically, the Taylor expansion $$f(z,\overline{z})=\underset{I,J=0}{\overset{\mathrm{}}{}}\frac{1}{I!J!}\frac{^{|I|+|J|}f}{z^I\overline{z}^J}(p)(zp)^I(\overline{z}\overline{p})^J$$ (3.12) of $`f𝒜_{p,\mathrm{}}C^\omega (^n)`$ converges unconditionally to $`f`$ with respect to the topology of $`𝒜_{p,\mathrm{}}`$. In particular, the truncated Taylor polynomials $$f^{(N,M)}(z,\overline{z})=\underset{I=0}{\overset{N}{}}\underset{J=0}{\overset{M}{}}\frac{1}{I!J!}\frac{^{|I|+|J|}f}{z^I\overline{z}^J}(p)(zp)^I(\overline{z}\overline{p})^J$$ (3.13) converge unconditionally to $`f`$ in the topology of $`𝒜_{p,\mathrm{}}`$. ###### Proof. Clearly, we only have to show the later statement. First we rewrite the seminorms $`_{m,\mathrm{},R,S}^{p,\mathrm{}}`$ in the following ‘measure-theoretic’ way $$h_{m,\mathrm{},R,S}^{p,\mathrm{}}\left(f\right)=\underset{I_1,J_1=0}{\overset{\mathrm{}}{}}\mathrm{}\underset{I_s,J_s=0}{\overset{\mathrm{}}{}}\mu _{I_1,\mathrm{}I_s,J_1,\mathrm{}J_s}^{m,\mathrm{},R,S,\mathrm{}}\left|\frac{^{|I_1|+|J_1|}f}{z^{I_1}\overline{z}^{J_1}}(p)\right|^2\mathrm{}\left|\frac{^{|I_s|+|J_s|}f}{z^{I_s}\overline{z}^{J_s}}(p)\right|^2,$$ where $`\mu _{I_1,\mathrm{}I_s,J_1,\mathrm{},J_s}^{m,\mathrm{},R,S,\mathrm{}}0`$ are numerical constants not depending on $`p`$, and $`s=2^m`$. This can be seen by induction easily. The concrete form of the coefficients $`\mu _{I_1,\mathrm{},I_s,J_1,\mathrm{},J_s}^{m,\mathrm{},R,S,\mathrm{}}`$ is not important for the following argument. Now we define for $`fC^{\mathrm{}}(^n)`$ a non-negative function $`\varphi _p(f):^{2sn}[0,\mathrm{})`$ by $$\varphi _p(f)(I_1,\mathrm{},I_s,J_1,\mathrm{},J_s)=\left|\frac{^{|I_1|+|J_1|}f}{z^{I_1}\overline{z}^{J_1}}(p)\right|^2\mathrm{}\left|\frac{^{|I_s|+|J_s|}f}{z^{I_s}\overline{z}^{J_s}}(p)\right|^2.$$ Then we can interpret $`h_{m,\mathrm{},R,S}^{p,\mathrm{}}\left(f\right)`$ as the ‘integral’ of $`\varphi _p(f)`$ over $`^{2sn}`$ with respect to the weighted counting measure $`\mathrm{d}\mu ^{m,\mathrm{},R,S,\mathrm{}}`$ determined by the coefficients $`\mu _{I_1,\mathrm{},I_s,J_1,\mathrm{},J_s}^{m,\mathrm{},R,S,\mathrm{}}`$, i.e. $$h_{m,\mathrm{},R,S}^{p,\mathrm{}}\left(f\right)=_{^{2sn}}\varphi _p(f)\mathrm{d}\mu ^{m,\mathrm{},R,S,\mathrm{}}.$$ Now let $`𝖪,𝖫^n`$ be finite subsets and define the polynomial $$f^{(𝖪,𝖫)}(z,\overline{z})=\underset{I𝖪}{}\underset{J𝖩}{}\frac{1}{I!J!}\frac{^{|I|+|J|f}}{z^I\overline{z}^J}(p)(zp)^I(\overline{z}\overline{p})^J.$$ Then we clearly have $$\varphi _p\left(ff^{(𝖪,𝖫)}\right)(I_1,\mathrm{},I_s,J_1,\mathrm{},J_s)=\{\begin{array}{cc}0\hfill & \text{if}I_1,\mathrm{},I_s𝖪,J_1,\mathrm{},J_s𝖫\hfill \\ \varphi _p(f)\hfill & \text{else},\hfill \end{array}$$ (3.14) Thus when $`𝖪,𝖫`$ exhaust $`^n`$, the function $`\varphi _p(ff^{(𝖪,𝖫)})`$ converges pointwise and monotonically to zero, i.e. for $`𝖪𝖪^{}`$ and $`𝖫𝖫^{}`$ we have $$\varphi _p\left(ff^{(𝖪,𝖫)}\right)\varphi _p\left(ff^{(𝖪^{},𝖫^{})}\right),$$ and for all $`I_1,\mathrm{},I_s,J_1,\mathrm{},J_s`$ $$\underset{𝖪,𝖫^n}{lim}\varphi _p\left(ff^{(𝖪,𝖫)}\right)(I_1,\mathrm{},I_s,J_1,\mathrm{},J_s)=0$$ in these sense of net convergence for the net of finite subsets of $`^n`$. Now an order of summation in (3.12) corresponds to a strictly increasing sequence $`𝖪_i\times 𝖫_i^n\times ^n`$ which exhausts $`^n\times ^n`$. Then $$\underset{i\mathrm{}}{lim}h_{m,\mathrm{},R,S}^{p,\mathrm{}}\left(ff^{(𝖪_i,𝖫_𝗂)}\right)=\underset{i\mathrm{}}{lim}\varphi _p\left(ff^{(𝖪_i,𝖫_𝗂)}\right)\mathrm{d}\mu ^{m,\mathrm{},R,S,\mathrm{}}=0$$ by dominated convergence. But this is equivalent to the unconditional convergence of (3.12), see also \[29, Sect. 14.6, Thm. 1\]. ∎ We now come to the main result of this section: ###### Theorem 3.10 The locally convex topology of $`𝒜_{p,\mathrm{}}`$ is complete, i.e. $`𝒜_{p,\mathrm{}}`$ is a Fréchet space. ###### Proof. Since the topology is determined by countably many seminorms we only have to consider Cauchy sequences and not Cauchy nets. Thus let $`f_i𝒜_{p,\mathrm{}}`$ be a Cauchy sequence, i.e. for all seminorms $`_{m,\mathrm{},R,S}^{p,\mathrm{}}`$ and all $`ϵ>0`$ we find a $`K(m,\mathrm{},R,S,ϵ)`$ such that for $`i,jK`$ we have $$f_if_j_{m,\mathrm{},R,S}^{p,\mathrm{}}<ϵ.$$ We first evaluate this for $`m=1,\mathrm{}=1,R,S=0`$. Let $$f_i(z,\overline{z})=\underset{I,J=0}{\overset{\mathrm{}}{}}\frac{1}{I!J!}a_{IJ}^{(i)}(zp)^I(\overline{z}\overline{p})^J$$ be the Taylor expansion of $`f_i`$ then from (3.11) we see that the Taylor coefficients $`a_{IJ}^{(i)}`$ form a Cauchy sequence for each $`I,J`$. Denote their limit by $$a_{IJ}=\underset{i\mathrm{}}{lim}a_{IJ}^{(i)}$$ and define $$f(z,\overline{z})=\underset{I,J=0}{\overset{\mathrm{}}{}}\frac{1}{I!J!}a_{IJ}(zp)^I(\overline{z}\overline{p})^J.$$ Then we want to show $`f𝒜_{p,\mathrm{}}`$ and $`f_if`$. To this end we first choose a smooth function $`\stackrel{~}{f}C^{\mathrm{}}(^n)`$ with Taylor coefficients at $`p`$ given by ($``$), which is possible thanks to the Borel Lemma. Since $`f_i`$ is a Cauchy sequence with respect to $`_{m,\mathrm{},R,S}^{p,\mathrm{}}`$ the sequence of seminorms $`f_i_{m,\mathrm{},R,S}^{p,\mathrm{}}`$ stays bounded as $`i0`$. Thus we can again use the measure-theoretic point of view and write with the notation from the previous proof $$\begin{array}{cc}\hfill h_{m,\mathrm{},R,S}^{p,\mathrm{}}\left(\stackrel{~}{f}\right)& =_{^{2sn}}\varphi (\stackrel{~}{f})\mathrm{d}\mu ^{m,\mathrm{},R,S,\mathrm{}}\hfill \\ & =_{^{2sn}}\underset{i\mathrm{}}{lim}\varphi (f_i)\mathrm{d}\mu ^{m,\mathrm{},R,S,\mathrm{}}\hfill \\ & =_{^{2sn}}\underset{i}{lim\; inf}\varphi (f_i)\mathrm{d}\mu ^{m,\mathrm{},R,S,\mathrm{}}\hfill \\ & \underset{i}{lim\; inf}_{^{2sn}}\varphi (f_i)\mathrm{d}\mu ^{m,\mathrm{},R,S,\mathrm{}}\hfill \\ & \underset{i}{sup}(f_i_{m,\mathrm{},R,S}^{p,\mathrm{}})^{2^{m+1}}<\mathrm{},\hfill \end{array}$$ by Fatou’s Lemma. Thus $`\stackrel{~}{f}\stackrel{~}{𝒜}_{p,\mathrm{}}`$ and hence $`f`$ as in ($``$) is the unique real-analytic representative in $`𝒜_{p,\mathrm{}}`$. Next we compute using (3.13) $$\begin{array}{cc}& ff_i_{m,\mathrm{},R,S}^{p,\mathrm{}}\hfill \\ & (ff_i)(ff_i)^{(N,M)}_{m,\mathrm{},R,S}^{p,\mathrm{}}+f^{(N,M)}f_j^{(N,M)}_{m,\mathrm{},R,S}^{p,\mathrm{}}+f_j^{(N,M)}f_i^{(N,M)}_{m,\mathrm{},R,S}^{p,\mathrm{}}\hfill \\ & (ff_i)(ff_i)^{(N,M)}_{m,\mathrm{},R,S}^{p,\mathrm{}}+f^{(N,M)}f_j^{(N,M)}_{m,\mathrm{},R,S}^{p,\mathrm{}}+f_jf_i_{m,\mathrm{},R,S}^{p,\mathrm{}}.\hfill \end{array}$$ Now we fix $`ϵ>0`$ and $`K`$ such that the last term is smaller than $`ϵ/3`$ for $`i,j>K`$. For such an $`i`$ we fix $`N,M`$ such that the first term is smaller $`ϵ/3`$ thanks to Theorem 3.9. Finally, for this choice of $`N,M`$ we can find $`j`$ large enough that the second term is smaller $`ϵ/3`$ since we have two polynomials of *fixed* degree $`(N,M)`$ whose coefficients converge. This finally proves $`f_if`$ with respect to $`_{m,\mathrm{},R,S}^{p,\mathrm{}}`$. ∎ ## 4 The continuity of $`_{\mathrm{Wick}}`$ From Proposition 3.7 we know that on the subspace $`[z,\overline{z}]𝒜_{p,\mathrm{}}`$ the Wick star product $`_{\mathrm{Wick}}^{\mathrm{}}`$ is well-defined and continuous with respect to the topology of $`𝒜_{p,\mathrm{}}`$. Since on the other hand $`[z,\overline{z}]`$ is a dense subspace by Theorem 3.9 the Wick star product extends uniquely to a continuous product on $`𝒜_{p,\mathrm{}}`$ which thereby becomes a Fréchet algebra. However, from this abstract extension we cannot yet conclude whether the *formula* (2.3) with $`\lambda `$ being replaced by $`\mathrm{}`$ is still true. Thus we need an additional argument. Let $`\mathrm{}\{0,\mathrm{},2^m1\}`$ be written as $`\mathrm{}=\mathrm{}_{m1}2^{m1}+\mathrm{}+\mathrm{}_12+\mathrm{}_0`$ with $`\mathrm{}_{m1},\mathrm{},\mathrm{}_0\{0,1\}`$. Then we define $`ϵ_{\mathrm{}}=(1)^{\mathrm{}_{m1}+\mathrm{}+\mathrm{}_0}`$. With this notation we can prove the following continuity property of the partial derivatives: ###### Proposition 4.1 Let $`fC^{\mathrm{}}(^n)`$ then we have $$\sqrt{(2\mathrm{})^{|I|+|J|}}\frac{^{|I|+|J|}f}{z^I\overline{z}^J}_{m,\mathrm{},R,S}^{p,\mathrm{}}\{\begin{array}{cc}f_{m,\mathrm{},R+I,S+J}^{p,\mathrm{}}\hfill & \text{for}ϵ_{\mathrm{}}=+1\hfill \\ f_{m,\mathrm{},R+J,S+I}^{p,\mathrm{}}\hfill & \text{for}ϵ_{\mathrm{}}=1.\hfill \end{array}$$ (4.1) ###### Proof. Clearly, for $`m=0`$ (and hence $`\mathrm{}=0`$) we even have the equality $$\sqrt{(2\mathrm{})^{|I|+|J|}}\frac{^{|I|+|J|}f}{z^I\overline{z}^J}_{0,0,R,S}^{p,\mathrm{}}=f_{0,0,R+I,S+J}^{p,\mathrm{}}$$ by the very definition of $`_{0,0,R,S}^{p,\mathrm{}}`$. Thus we prove the claim by induction on $`m`$. Let first be $`\mathrm{}`$ even and $`ϵ_{\mathrm{}}=+1`$. Then $`ϵ_{\mathrm{}/2}=+1`$ as well and we have by induction $$\begin{array}{cc}\hfill h_{m,\mathrm{},R,S}^{p,\mathrm{}}\left(\sqrt{(2\mathrm{})^{|I|+|J|}}\frac{^{|I|+|J|}f}{z^I\overline{z}^J}\right)& \underset{N=0}{\overset{\mathrm{}}{}}\frac{1}{N!}\left(\underset{K=0}{\overset{R}{}}\underset{L=0}{\overset{N+S}{}}\left(\genfrac{}{}{0.0pt}{}{R}{K}\right)\left(\genfrac{}{}{0.0pt}{}{N+S}{L}\right)h_{m1,\mathrm{}/2,K+I,L+J}^{p,\mathrm{}}\left(f\right)\right)^2\hfill \\ & =\underset{N=0}{\overset{\mathrm{}}{}}\frac{1}{N!}\left(\underset{K=I}{\overset{R+I}{}}\underset{L=J}{\overset{N+S+J}{}}\left(\genfrac{}{}{0.0pt}{}{R}{KI}\right)\left(\genfrac{}{}{0.0pt}{}{N+S}{LJ}\right)h_{m1,\mathrm{}/2,K,L}^{p,\mathrm{}}\left(f\right)\right)^2\hfill \\ & \underset{N=0}{\overset{\mathrm{}}{}}\frac{1}{N!}\left(\underset{K=I}{\overset{R+I}{}}\underset{L=J}{\overset{N+S+J}{}}\left(\genfrac{}{}{0.0pt}{}{R+I}{K}\right)\left(\genfrac{}{}{0.0pt}{}{N+S+J}{L}\right)h_{m1,\mathrm{}/2,K,L}^{p,\mathrm{}}\left(f\right)\right)^2\hfill \\ & h_{m,\mathrm{},R+I,S+J}^{p,\mathrm{}}\left(f\right).\hfill \end{array}$$ For $`ϵ_{\mathrm{}}=ϵ_{\mathrm{}/2}=1`$ we get $`h_{m,\mathrm{},R+J,S+I}^{p,\mathrm{}}\left(f\right)`$ instead and the two cases with odd $`\mathrm{}`$ are analogous. ∎ From Theorem 3.9 we see that the polynomials $$\zeta _p^{IJ}(z,\overline{z})=(zp)^I(\overline{z}\overline{p})^J$$ (4.2) form a countable unconditional topological basis for $`𝒜_{p,\mathrm{}}`$. The proposition now implies that we even have a *Schauder basis*: ###### Corollary 4.2 The polynomials $`\{\zeta _p^{IJ}\}_{I,J^n}`$ form an unconditional Schauder basis for $`𝒜_{p,\mathrm{}}`$. ###### Proof. By Theorem 3.9 we have the unconditional convergence $$f=\underset{I,J=0}{\overset{\mathrm{}}{}}\frac{1}{I!J!}\delta _p^{IJ}(f)\zeta _p^{IJ},$$ and the $`(I,J)`$-th derivative $`\delta _p^{IJ}`$ of the $`\delta `$-functionals are continuous linear functionals on $`𝒜_{p,\mathrm{}}`$ by Proposition 4.1. This implies the result, see e.g. \[29, Sect. 14.2\] for a definition of a Schauder basis. ∎ The next proposition shows that the pointwise product is continuous in the topology of $`𝒜_{p,\mathrm{}}`$: ###### Proposition 4.3 Let $`f,gC^{\mathrm{}}(^n)`$. Then we have $$fg_{m,\mathrm{},R,S}^{p,\mathrm{}}f_{m+1,\mathrm{},R,S}^{p,\mathrm{}}g_{m+1,\mathrm{},R,S}^{p,\mathrm{}},$$ (4.3) whence $`𝒜_{p,\mathrm{}}`$ is a Fréchet -algebra with respect to the pointwise product. ###### Proof. First recall that from the explicit form of the Wick star product we obtain $$(2\mathrm{})^{|I|+|J|}\left|\frac{^{|I|+|J|}f}{z^I\overline{z}^J}\right|^2(p)h_{0,0,I,J}^{p,\mathrm{}}\left(f\right).$$ Now we first consider the case $`m=0`$. Here we have by the Leibniz rule $$\begin{array}{cc}& h_{0,0,R,S}^{p,\mathrm{}}\left(fg\right)\hfill \\ & =\underset{N=0}{\overset{\mathrm{}}{}}\frac{(2\mathrm{})^{|N|+|R|+|S|}}{N!}\left|\underset{I=0}{\overset{R}{}}\underset{J=0}{\overset{N+S}{}}\left(\genfrac{}{}{0.0pt}{}{R}{I}\right)\left(\genfrac{}{}{0.0pt}{}{N+S}{J}\right)\frac{^{|I|+|J|}f}{z^I\overline{z}^J}\frac{^{|RI|+|N+SJ|}g}{z^{RI}\overline{z}^{N+SJ}}\right|^2(p)\hfill \\ & \underset{N=0}{\overset{\mathrm{}}{}}\frac{(2\mathrm{})^{|N|+|R|+|S|}}{N!}\left(\underset{I=0}{\overset{R}{}}\underset{J=0}{\overset{N+S}{}}\left(\genfrac{}{}{0.0pt}{}{R}{I}\right)\left(\genfrac{}{}{0.0pt}{}{N+S}{J}\right)\left|\frac{^{|I|+|J|}f}{z^I\overline{z}^J}(p)\right|\left|\frac{^{|RI|+|N+SJ|}g}{z^{RI}\overline{z}^{N+SJ}}(p)\right|\right)^2\hfill \\ & \stackrel{()}{}\underset{N=0}{\overset{\mathrm{}}{}}\frac{1}{N!}\left(\underset{I=0}{\overset{R}{}}\underset{J=0}{\overset{N+S}{}}\left(\genfrac{}{}{0.0pt}{}{R}{I}\right)\left(\genfrac{}{}{0.0pt}{}{N+S}{J}\right)\sqrt{h_{0,0,I,J}^{p,\mathrm{}}\left(f\right)}\sqrt{h_{0,0,RI,N+SJ}^{p,\mathrm{}}\left(g\right)}\right)^2\hfill \\ & \underset{N=0}{\overset{\mathrm{}}{}}\frac{1}{N!}\left(\underset{I=0}{\overset{R}{}}\underset{J=0}{\overset{N+S}{}}\left(\genfrac{}{}{0.0pt}{}{R}{I}\right)\left(\genfrac{}{}{0.0pt}{}{N+S}{J}\right)h_{0,0,I,J}^{p,\mathrm{}}\left(f\right)\right)\left(\underset{I=0}{\overset{R}{}}\underset{J=0}{\overset{N+S}{}}\left(\genfrac{}{}{0.0pt}{}{R}{I}\right)\left(\genfrac{}{}{0.0pt}{}{N+S}{J}\right)h_{0,0,I,J}^{p,\mathrm{}}\left(g\right)\right)\hfill \\ & \sqrt{\underset{N=0}{\overset{\mathrm{}}{}}\frac{1}{N!}\left(\underset{I=0}{\overset{R}{}}\underset{J=0}{\overset{N+S}{}}\left(\genfrac{}{}{0.0pt}{}{R}{I}\right)\left(\genfrac{}{}{0.0pt}{}{N+S}{J}\right)h_{0,0,I,J}^{p,\mathrm{}}\left(f\right)\right)^2}\hfill \\ & \times \sqrt{\underset{N=0}{\overset{\mathrm{}}{}}\frac{1}{N!}\left(\underset{I=0}{\overset{R}{}}\underset{J=0}{\overset{N+S}{}}\left(\genfrac{}{}{0.0pt}{}{R}{I}\right)\left(\genfrac{}{}{0.0pt}{}{N+S}{J}\right)h_{0,0,I,J}^{p,\mathrm{}}\left(g\right)\right)^2}\hfill \\ & =\sqrt{h_{1,0,R,S}^{p,\mathrm{}}\left(f\right)}\sqrt{h_{1,0,R,S}^{p,\mathrm{}}\left(g\right)},\hfill \end{array}$$ using twice the Cauchy Schwarz inequality in the last steps. For $`m1`$ we proceed by a straightforward induction for the two cases of $`\mathrm{}`$ even and odd separately. ∎ ###### Corollary 4.4 $`𝒜_{p,\mathrm{}}`$ is a Fréchet-Poisson -algebra. ###### Proof. Clearly, the canonical Poisson bracket is continuous as $`\{f,g\}`$ is the sum of pointwise products of partial derivatives of $`f`$ and $`g`$. ∎ Combining the last two propositions we can finally show the continuity of the Wick star product: we even show that the series (2.3) converges in the topology of $`𝒜_{p,\mathrm{}}`$ if we replace $`\lambda `$ by any complex number $`\alpha `$: ###### Theorem 4.5 The Wick star product $$f_{\mathrm{Wick}}^\alpha g=\underset{N=0}{\overset{\mathrm{}}{}}\frac{(2\alpha )^{|N|}}{N!}\frac{^{|N|}f}{z^N}\frac{^{|N|}g}{\overline{z}^N}$$ (4.3) converges absolutely in the topology of $`𝒜_{p,\mathrm{}}`$ for all $`\alpha `$ and gives a continuous associative product. If $`\alpha =\mathrm{}>0`$ then $`𝒜_{p,\mathrm{}}`$ becomes a Fréchet -algebra with respect to $`_{\mathrm{Wick}}^{\mathrm{}}`$ as the product and the complex conjugation as the -involution. ###### Proof. Let $`f,g𝒜_{p,\mathrm{}}`$ then we have for even $`\mathrm{}`$ and $`ϵ_{\mathrm{}}=+1`$ $$\begin{array}{cc}& \underset{N=0}{\overset{\mathrm{}}{}}\frac{(2\alpha )^{|N|}}{N!}\frac{^{|N|}f}{z^N}\frac{^{|N|}g}{\overline{z}^N}_{m,\mathrm{},0,0}^{p,\mathrm{}}\hfill \\ & \underset{N=0}{\overset{\mathrm{}}{}}\frac{|2\alpha |^{|N|}}{N!}\frac{^{|N|}f}{z^N}\frac{^{|N|}g}{\overline{z}^N}_{m,\mathrm{},0,0}^{p,\mathrm{}}\hfill \\ & \stackrel{\text{Prop. }\text{4.3}}{}\underset{N=0}{\overset{\mathrm{}}{}}\frac{|2\alpha |^{|N|}}{N!}\frac{^{|N|}f}{z^N}_{m+1,\mathrm{},0,0}^{p,\mathrm{}}\frac{^{|N|}g}{\overline{z}^N}_{m+1,\mathrm{},0,0}^{p,\mathrm{}}\hfill \\ & \stackrel{\text{Prop. }\text{4.1}}{}\underset{N=0}{\overset{\mathrm{}}{}}\frac{|\frac{\alpha }{\mathrm{}}|^{|N|}}{N!}f_{m+1,\mathrm{},N,0}^{p,\mathrm{}}g_{m+1,\mathrm{},0,N}^{p,\mathrm{}}\hfill \\ & \sqrt{\underset{N=0}{\overset{\mathrm{}}{}}\frac{|\frac{\alpha }{\mathrm{}}|^{|N|}}{N!}\left(f_{m+1,\mathrm{},N,0}^{p,\mathrm{}}\right)^2}\sqrt{\underset{N=0}{\overset{\mathrm{}}{}}\frac{|\frac{\alpha }{\mathrm{}}|^{|N|}}{N!}\left(g_{m+1,\mathrm{},0,N}^{p,\mathrm{}}\right)^2}\hfill \\ & \stackrel{\text{Prop. }\text{3.3}}{}\sqrt{\underset{N=0}{\overset{\mathrm{}}{}}\frac{|\frac{\alpha }{\mathrm{}}|^{|N|}}{N!}\left(\sqrt[2^{m+3}]{N!}f_{m+2,2\mathrm{}+1,0,0}^{p,\mathrm{}}\right)^2}\sqrt{\underset{N=0}{\overset{\mathrm{}}{}}\frac{|\frac{\alpha }{\mathrm{}}|^{|N|}}{N!}\left(\sqrt[2^{m+3}]{N!}g_{m+2,2\mathrm{},0,0}^{p,\mathrm{}}\right)^2}\hfill \\ & =\underset{=c_m(\frac{\alpha }{\mathrm{}})}{\underset{}{\left(\underset{N=0}{\overset{\mathrm{}}{}}\frac{|\frac{\alpha }{\mathrm{}}|^{|N|}}{N!}\sqrt[2^{m+2}]{N!}\right)}}f_{m+2,2\mathrm{}+1,0,0}^{p,\mathrm{}}g_{m+2,2\mathrm{},0,0}^{p,\mathrm{}}.\hfill \end{array}$$ Since $`c_m(\frac{\alpha }{\mathrm{}})`$ converges for all $`\alpha `$ we have shown the convergence of (4.3) with respect to $`_{m,\mathrm{},0,0}^{p,\mathrm{}}`$ for even $`\mathrm{}`$ and $`ϵ_{\mathrm{}}=+1`$. The other three cases are shown analogously. As the topology of $`𝒜_{p,\mathrm{}}`$ is already determined by the seminorms $`_{m,\mathrm{},0,0}^{p,\mathrm{}}`$ the *absolute* convergence in the topology of $`𝒜_{p,\mathrm{}}`$ follows. From the above estimate (and the analogous ones for odd $`\mathrm{}`$ etc.) one also obtains the continuity of $`_{\mathrm{Wick}}^\alpha `$. If $`\alpha =\mathrm{}`$ is real, then the complex conjugation is a -involution showing the last statement. ∎ ###### Corollary 4.6 The Wick star product $`_{\mathrm{Wick}}^\alpha `$ is a holomorphic deformation of the pointwise product in the sense of . ###### Corollary 4.7 For $`f,g𝒜_{p,\mathrm{}}`$ we have $$\overline{f}_{\mathrm{Wick}}^{\mathrm{}}g_{m,\mathrm{},R,S}^{p,\mathrm{}}f_{m+1,2^m+\mathrm{},R,S}^{p,\mathrm{}}g_{m+1,\mathrm{},R,S}^{p,\mathrm{}}.$$ (4.4) ###### Proof. This follows from Proposition 3.7, the density of $`[z,\overline{z}]`$ in $`𝒜_{p,\mathrm{}}`$ and the continuity of $`_{\mathrm{Wick}}^{\mathrm{}}`$. ∎ Though $`𝒜_{p,\mathrm{}}`$ becomes a Fréchet -algebra, the topology is *not* locally $`m`$-convex in the sense of , see also \[43, App. A\]. Recall that a locally convex algebra is called locally $`m`$-convex if there exist a set of seminorms $`_i`$ defining the topology such that $`ab_ia_ib_i`$. Such locally $`m`$-convex algebras always have a holomorphic calculus which fails for $`𝒜_{p,\mathrm{}}`$: ###### Example 4.8 We consider the entire function $`f𝒪()`$ defined by $$f(z)=\underset{r=0}{\overset{\mathrm{}}{}}\frac{z^r}{\sqrt[4]{r!}}.$$ (4.5) Then it is easy to see that $`f_{\mathrm{Wick}}^{\mathrm{}}\overline{f}`$ evaluated at $`z=0`$ converges only for $`\mathrm{}=0`$. Since clearly the $`\delta `$-functional $`\delta _p:𝒜_{p,\mathrm{}}`$ is continuous we conclude that $`f𝒜_{0,\mathrm{}}`$. This shows that $`𝒜_{0,\mathrm{}}`$ does *not* allow a holomorphic functional calculus as $`z𝒜_{0,\mathrm{}}`$ and the $`_{\mathrm{Wick}}^{\mathrm{}}`$-Taylor expansion of $`f`$ would again coincide with $`f`$ since $`_{\mathrm{Wick}}^{\mathrm{}}`$-power of $`z`$ coincide with the corresponding pointwise powers. Analogous arguments apply also for $`p0`$ and higher dimensions $`n1`$. ###### Corollary 4.9 The topology of $`𝒜_{p,\mathrm{}}`$ is not locally $`m`$-convex with respect to the Wick star product $`_{\mathrm{Wick}}^\alpha `$ for all $`\alpha `$ and there is no general holomorphic calculus for $`𝒜_{p,\mathrm{}}`$. In the following we shall equip $`𝒜_{p,\mathrm{}}`$ always with the Wick star product $`_{\mathrm{Wick}}^{\mathrm{}}`$. ###### Remark 4.10 At this point it would be interesting to compare our algebra to the construction obtained in : Here the authors consider the Weyl-Moyal star product, which on the formal level is known to be equivalent to the Wick star product, and establish a convergence scheme to obtain a certain Fréchet algebra as the completion of the polynomials. However, their construction is rather different from ours whence it seems difficult to investigate whether the usual formal equivalence transformation survives the convergence conditions. ## 5 Translations and rescalings We shall now discuss the dependence of $`𝒜_{p,\mathrm{}}`$ on the point $`p^n`$ and on the value $`\mathrm{}>0`$. We start with the dependence on the point $`p`$. Let $`\alpha ^n`$ then for $`f𝒜_{p,\mathrm{}}`$ we define $$(\tau _\alpha f)(z,\overline{z})=\widehat{f}(z+\alpha ,\overline{z})$$ (5.1) and $$(\overline{\tau }_{\overline{\alpha }}f)(z,\overline{z})=\widehat{f}(z,\overline{z}+\overline{\alpha }),$$ (5.2) which is well-defined according to Theorem 3.6, Part iii.). Moreover, we consider the functions $$𝖾_{\overline{\alpha },\beta }(z,\overline{z})=\mathrm{e}^{\mathrm{}\overline{\alpha }\beta }\mathrm{e}^{\overline{\alpha }z+\beta \overline{z}},$$ (5.3) which are elements in $`𝒜_{p,\mathrm{}}`$ according to Theorem 3.6, Part iv.). The following lemma is a simple computation, the results of which are well-known in the case of the formal Wick star product: ###### Lemma 5.1 Let $`\alpha ,\beta ,\gamma ,\delta ^n`$. Then we have for all $`f𝒜_{p,\mathrm{}}`$: 1. $`𝖾_{\overline{\alpha },\beta }_{\mathrm{Wick}}^{\mathrm{}}𝖾_{\overline{\gamma },\delta }=\mathrm{e}^{\mathrm{}(\overline{\alpha }\delta \beta \overline{\gamma })}𝖾_{\overline{\alpha }+\overline{\gamma },\beta +\delta }`$. 2. $`𝖾_{\overline{\alpha },\beta }_{\mathrm{Wick}}^{\mathrm{}}f=𝖾_{\overline{\alpha },\beta }\overline{\tau }_{2\mathrm{}\overline{\alpha }}f`$. 3. $`f_{\mathrm{Wick}}^{\mathrm{}}𝖾_{\overline{\alpha },\beta }=𝖾_{\overline{\alpha },\beta }\tau _{2\mathrm{}\beta }f`$. 4. The maps $`\tau _\alpha `$ and $`\overline{\tau }_{\overline{\alpha }}`$ are continuous linear bijections $$\tau _\alpha ,\overline{\tau }_{\overline{\alpha }}:𝒜_{p,\mathrm{}}𝒜_{p,\mathrm{}}.$$ (5.4) ###### Proof. The only non-trivial point here is that for $`f𝒜_{p,\mathrm{}}`$ we have $$\tau _\alpha f=\underset{N=0}{\overset{\mathrm{}}{}}\frac{\alpha ^N}{N!}\frac{^|N|f}{z^N}$$ and analogously for $`\overline{\tau }_{\overline{\alpha }}`$ since $`f`$ has a extension to $`\widehat{f}𝒪\times \overline{𝒪}(^n\times ^n)`$. Then the computations for the first three parts are folklore. The last part follows from $$\tau _\alpha f=𝖾_{0,\frac{\alpha }{2\mathrm{}}}\left(f_{\mathrm{Wick}}^{\mathrm{}}𝖾_{0,\frac{\alpha }{2\mathrm{}}}\right)$$ and the continuity of the pointwise multiplication as well as of the continuity of $`_{\mathrm{Wick}}^{\mathrm{}}`$. The same argument applies for $`\overline{\tau }_{\overline{\alpha }}`$. From ($``$) one can easily work out explicit estimates for $`\tau _\alpha f_{m,\mathrm{},R,S}^{p,\mathrm{}}`$ and $`\overline{\tau }_{\overline{\alpha }}f_{m,\mathrm{},R,S}^{p,\mathrm{}}`$ using Proposition 4.3 and Corollary 4.7. ∎ ###### Corollary 5.2 Let $`\alpha ,\beta ^n`$. 1. $`𝖾_{\overline{\alpha },\beta }𝒜_{p,\mathrm{}}`$ is invertible with respect to $`_{\mathrm{Wick}}^{\mathrm{}}`$ with inverse given by $`𝖾_{\overline{\alpha },\beta }`$. 2. $`𝖾_{\overline{\alpha },\beta }`$ is unitary iff $`\alpha =\beta `$ since in general $`\overline{𝖾_{\overline{\alpha },\beta }}=𝖾_{\overline{\beta },\alpha }`$. We introduce now the following notation. For $`w^n`$ we denote the translation by $`w`$ by $`T_w(z)=z+w`$ whence we have the corresponding pull-back on functions $`(T_w^{}f)(z)=f(z+w)`$. Moreover, we set $$u_w=𝖾_{\frac{1}{2\mathrm{}}\overline{w},\frac{1}{2\mathrm{}}w}𝒜_{p,\mathrm{}},$$ (5.4) which is a unitary element of $`𝒜_{p,\mathrm{}}`$ according to Corollary 5.2. ###### Proposition 5.3 The translation group $`^n`$ acts via pull-backs by continuous inner -automorphisms $$T_w^{}=\mathrm{Ad}_{\mathrm{Wick}}(u_w)$$ (5.5) on $`𝒜_{p,\mathrm{}}`$. ###### Proof. This is a simple consequence of Lemma 5.1 and the fact that $`u_w`$ is unitary. ∎ ###### Remark 5.4 We remark that on a heuristic level (or on polynomial functions only) the statement of this proposition is folklore. Note that it is clear, that for the *formal* Wick star product the statement is *wrong*: the translations are only *outer* automorphisms as the elements $`u_w`$ are *not* well-defined as formal series in $`\lambda `$. In fact, it was one of our main motivations to find a reasonably large algebra where the statement of the proposition is still true, extending the polynomials. The fact that the translations act by inner automorphisms will immediately imply the following result: ###### Theorem 5.5 Let $`p,p^{}^n`$. Then $$𝒜_{p,\mathrm{}}=𝒜_{p^{},\mathrm{}}$$ (5.6) as Fréchet -algebras. ###### Proof. Let $`f𝒜_{p,\mathrm{}}`$ be given and $`w=p^{}p`$. Then we have on one hand $$f_{m,\mathrm{},R,S}^{p^{},\mathrm{}}=T_w^{}f_{m,\mathrm{},R,S}^{p,\mathrm{}}<\mathrm{}$$ since $$\frac{^{|I|+|J|}f}{z^I\overline{z}^J}(p^{})=\frac{^{|I|+|J|}(T_w^{}f)}{z^I\overline{z}^J}(p),$$ and in the definition of $`_{m,\mathrm{},R,S}^{p,\mathrm{}}`$ only the Taylor coefficients of $`f`$ at $`p`$ are used while the combinatorial coefficients in the construction are the same for all points $`p^{}`$, $`p`$. This shows $`f𝒜_{p^{},\mathrm{}}`$ whence by symmetry the equality (5.6) as vector spaces follows. On the other hand we have $$f_{m,\mathrm{},R,S}^{p^{},\mathrm{}}=u_w_{\mathrm{Wick}}^{\mathrm{}}f_{\mathrm{Wick}}^{\mathrm{}}u_w_{m,\mathrm{},R,S}^{p,\mathrm{}}\overline{u_w}_{m+1,2^m+\mathrm{},R,S}^{p,\mathrm{}}f_{m+2,2^{m+1}+\mathrm{},R,S}^{p,\mathrm{}}u_w_{m+2,\mathrm{},R,S}^{p,\mathrm{}}.$$ This shows that the seminorms $`_{m,\mathrm{},R,S}^{p^{},\mathrm{}}`$ can be estimated against the seminorms $`_{m,\mathrm{},R,S}^{p,\mathrm{}}`$ whence the topology of $`𝒜_{p^{},\mathrm{}}`$ is coarser than the one of $`𝒜_{p,\mathrm{}}`$. By symmetry $`p^{}p`$ we see that they actually coincide. The algebraic structures are the same anyway whence the theorem is shown. ∎ Thus we see *a posteriori* that the construction of the Fréchet algebra $`𝒜_{p,\mathrm{}}`$ does not depend on our choice $`p^n`$. Hence we can simply write $`𝒜_{\mathrm{}}=𝒜_{p,\mathrm{}}`$ from now on. Note however, that the system of seminorms $`_{m,\mathrm{},R,S}^{p,\mathrm{}}`$ *depends* on the choice of $`p`$, only the induced topology is independent of $`p`$. As in the formal case, *all* $`\delta `$-functionals are positive: ###### Corollary 5.6 All $`\delta `$-functionals $$\delta _p:𝒜_{\mathrm{}}$$ (5.7) are continuous positive linear functionals. ###### Proof. The positivity is clear from (4.3) with $`\alpha =\mathrm{}>0`$ and the continuity follows from the theorem as $`|\delta _p(f)|f_{0,0,0,0}^{p,\mathrm{}}`$. ∎ ###### Corollary 5.7 The Fréchet topology of $`𝒜_{\mathrm{}}`$ is finer than the topology of pointwise convergence. In the next step we want to analyze the continuity properties of the group representation $`wT_w^{}`$ further. To this end we consider the dependence of the elements $`u_w`$ on $`w`$. Since $$u_w_{\mathrm{Wick}}^{\mathrm{}}u_v=\mathrm{e}^{\frac{\mathrm{i}}{2\mathrm{}}\mathrm{Im}(\overline{w}v)}u_{w+v}$$ (5.8) the map $`wu_w`$ is *not* a group morphism. Since $`\mathrm{e}^{\frac{\mathrm{i}}{2\mathrm{}}\mathrm{Im}(\overline{w}v)}`$ is even a non-trivial group cocycle we need to pass to the central extension of the translation group $`^n`$ by this cocycle, i.e. to the Heisenberg group $`𝖧_n`$. Here we use the convention that $`𝖧_n=^n\times `$ with multiplication law $$(w,c)(w^{},c^{})=(w+w^{},c+c^{}+\mathrm{Im}(\overline{w}w^{})).$$ (5.9) Then it follows that $$𝖧_n(w,c)u_{(w,c)}=\mathrm{e}^{\frac{\mathrm{i}}{2\mathrm{}}c}u_w\mathrm{U}(𝒜_{\mathrm{}})$$ (5.10) is a group morphism from $`𝖧_n`$ into the group of unitaries $`\mathrm{U}(𝒜_{\mathrm{}})`$ in $`𝒜_{\mathrm{}}`$. Since $`\mathrm{e}^{\frac{\mathrm{i}}{2\mathrm{}}c}`$ is central, (5.10) factors to the group morphism $`wT_w^{}`$, i.e. we have $$\mathrm{Ad}_{\mathrm{Wick}}(u_{(w,c)})=\mathrm{Ad}_{\mathrm{Wick}}(u_w)=T_w^{}$$ (5.11) for all $`(w,c)𝖧_n`$. The Lie algebra $`𝔥^n`$ of $`𝖧_n`$ can be identified with $`𝖧_n`$ via the exponential map. Then the Lie bracket is given by $`[(w,c),(w^{},c^{})]=(0,2\mathrm{Im}(\overline{w}w^{}))`$. The next theorem shows that the group morphism $`𝖧_n\mathrm{U}(𝒜_{\mathrm{}})`$ is analytic and induces a Lie algebra morphism $`𝔥^n𝒜_{\mathrm{}}`$: ###### Theorem 5.8 1. The map $`𝔥^n𝖧_n(w,c)u_{(w,c)}`$ is analytic with respect to the topology of $`𝒜_{\mathrm{}}`$. 2. The generator $`J_{(w,c)}`$ of the one-parameter group $`tu_{(tw,tc)}`$ is given by $$J_{(w,c)}(z,\overline{z})=\frac{\mathrm{d}}{\mathrm{d}t}|_{t=0}u_{(tw,tc)}=\frac{\mathrm{i}}{2\mathrm{}}c+\frac{1}{2\mathrm{}}(\overline{w}zw\overline{z}),$$ (5.12) and $`𝔥^n(w,c)J_{(w,c)}𝒜_{\mathrm{}}`$ is a Lie algebra morphism where $`𝒜_{\mathrm{}}`$ is equipped with the $`_{\mathrm{Wick}}^{\mathrm{}}`$-commutator as Lie bracket. 3. We have for all $`k`$ $$\frac{\mathrm{d}^k}{\mathrm{d}t^k}|_{t=0}u_{(tw,tc)}=J_{(w,c)}\underset{k\text{times}}{\underset{}{_{\mathrm{Wick}}^{\mathrm{}}\mathrm{}_{\mathrm{Wick}}^{\mathrm{}}}}J_{(w,c)},$$ (5.13) and explicitly for $`c=0`$ $$\frac{\mathrm{d}^k}{\mathrm{d}t^k}|_{t=0}u_{(tw,0)}=\underset{\mathrm{}=0}{\overset{k/2}{}}\frac{k!}{\mathrm{}!(k2\mathrm{})!}\left(\frac{\overline{w}w}{4\mathrm{}}\right)^{\mathrm{}}\left(\frac{\overline{w}zw\overline{z}}{2\mathrm{}}\right)^{k2\mathrm{}}.$$ (5.14) ###### Proof. We have $`u_{(w,c)}(z,\overline{z})=\mathrm{e}^{\frac{\mathrm{i}}{2\mathrm{}}c\frac{\overline{w}w}{4\mathrm{}}+\frac{1}{2\mathrm{}}(\overline{w}zw\overline{z})}`$. The first factors $`\mathrm{e}^{\frac{\mathrm{i}}{2\mathrm{}}c}`$ and $`\mathrm{e}^{\frac{\overline{w}w}{2\mathrm{}}}`$ are clearly analytic as they are analytic functions times a fixed element (the identity) in $`𝒜_{\mathrm{}}`$. For the remaining factor we see that the $`(z,\overline{z})`$-Taylor expansion, which converges unconditionally in the topology of $`𝒜_{\mathrm{}}`$ coincides up to numerical factors with the $`(w,\overline{w})`$-Taylor expansion of this function. Thus the $`(w,\overline{w})`$-Taylor expansion converges also in $`𝒜_{\mathrm{}}`$ unconditionally which shows the first part. The second part is a trivial computation. For the third part, the contribution of $`c`$ is not essential whence we discuss the case $`c=0`$ only. A straightforward computation of the Taylor expansion in $`t`$ gives immediately (5.14). On the other hand, the $`k`$-th power of the linear function $`J_w=J_{(w,0)}`$ satisfies the recursion formula $$J_w^k=J_wJ_w^{(k1)}+(k1)\frac{\overline{w}w}{2\mathrm{}}J_w^{(k2)},$$ which follows easily from the fact that partial derivatives are derivations of $`_{\mathrm{Wick}}^{\mathrm{}}`$ and $`\overline{w}\frac{}{\overline{z}}J_w=\frac{\overline{w}w}{2\mathrm{}}`$ is central. In a last step one shows that also the right hand side of (5.14) satisfies this recursion with the same initial conditions for $`k=0,1`$. ∎ ###### Corollary 5.9 Let $`w^n`$ then the $`_{\mathrm{Wick}}^{\mathrm{}}`$-exponential function $$\mathrm{Exp}\left(tJ_w\right)=\underset{k=0}{\overset{\mathrm{}}{}}\frac{t^k}{k!}J_w\underset{k\text{times}}{\underset{}{_{\mathrm{Wick}}^{\mathrm{}}\mathrm{}_{\mathrm{Wick}}^{\mathrm{}}}}J_w=u_{tw}$$ (5.15) converges unconditionally in the topology of $`𝒜_{\mathrm{}}`$ for all $`t`$. ###### Remark 5.10 Again, the importance of this corollary is not the explicit computation of the star exponential which is folklore. Instead, we have found a well-defined analytic framework where the formula actually converges inside an *algebra* of functions. Note also, that in general we cannot expect such a convergence as $`𝒜_{\mathrm{}}`$ does not allow a holomorphic functional calculus in general. Let us now discuss the dependence on the parameter $`\mathrm{}`$. From a simple dimensional analysis we see that with our convention for the Poisson bracket $`\{z^k,\overline{z}^{\mathrm{}}\}=\frac{2}{\mathrm{i}}\delta ^k\mathrm{}`$ the coordinates have to have the physical dimension $`[\text{action}]^{\frac{1}{2}}`$. Thus a rescaling of $`\mathrm{}`$, which physically is of course absurd, has to be reinterpreted as a rescaling of the coordinates $`(z,\overline{z})`$: we are not changing the value $`\mathrm{}`$ but the unit system. On the other hand, from a purely mathematical point of view we cannot distinguish these two interpretations. As we do not have any additional absolute scale in our approach, the corresponding algebras $`𝒜_{\mathrm{}}`$ and $`𝒜_{\mathrm{}^{}}`$ should be isomorphic in order to be physically reasonable. The following theorem will show that this is indeed the case. We define for $`\alpha >0`$ the diffeomorphism $`R_\alpha :^n^n`$ $$R_\alpha (z)=\sqrt{\alpha }z,$$ (5.16) whose inverse is $`R_{\frac{1}{\alpha }}`$. ###### Theorem 5.11 The pull-back $`R_\alpha ^{}`$ induces an isomorphism of Fréchet -algebras $$R_\alpha ^{}:𝒜_\alpha \mathrm{}𝒜_{\mathrm{}}.$$ (5.17) In particular, we have for all $`f𝒜_\alpha \mathrm{}`$ $$R_\alpha ^{}f_{m,\mathrm{},R,S}^{0,\mathrm{}}=f_{m,\mathrm{},R,S}^{0,\alpha \mathrm{}}.$$ (5.18) ###### Proof. Let $`f𝒜_\alpha \mathrm{}`$ be given. Then we have $$\frac{^{|I|+|J|}(R_\alpha ^{}f)}{z^I\overline{z}^J}=\sqrt{\alpha }^{|I|+|J|}R_\alpha ^{}\left(\frac{^{|I|+|J|}f}{z^I\overline{z}^J}\right)$$ and thus $$\begin{array}{cc}\hfill R_\alpha ^{}f_{0,0,R,S}^{0,\mathrm{}}& =\underset{N=0}{\overset{\mathrm{}}{}}\frac{(2\mathrm{})^{|N|+|R|+|S|}}{N!}\left|\frac{^{|N|+|R|+|S|}(R_\alpha ^{}f)}{z^R\overline{z}^{N+S}}(0)\right|^2\hfill \\ & =\underset{N=0}{\overset{\mathrm{}}{}}\frac{(2\alpha \mathrm{})^{|N|+|R|+|S|}}{N!}\left|\frac{^{|N|+|R|+|S|}f}{z^R\overline{z}^{N+S}}(0)\right|^2\hfill \\ & =f_{0,0,R,S}^{0,\alpha \mathrm{}}.\hfill \end{array}$$ Since the higher seminorms are constructed in a purely combinatorial way out of $`_{0,0,R,S}^{0,\mathrm{}}`$, we can conclude (5.18). This shows that $`R_\alpha ^{}:𝒜_\alpha \mathrm{}𝒜_{\mathrm{}}`$ is an isomorphism of Fréchet spaces as we can exchange the role of $`\mathrm{}`$ and $`\alpha \mathrm{}`$ by passing from $`\alpha `$ to $`\frac{1}{\alpha }`$. Then it is easy to see that $`R_\alpha ^{}`$ is an algebra morphism: we use the convergence of (4.3) and the continuity of $`R_\alpha ^{}`$ to obtain $$\begin{array}{cc}\hfill R_\alpha ^{}(f_{\mathrm{Wick}}^\alpha \mathrm{}g)& =R_\alpha ^{}\left(\underset{N=0}{\overset{\mathrm{}}{}}\frac{(2\alpha \mathrm{})^{|N|}}{N!}\frac{^{|N|}f}{z^N}\frac{^{|N|}g}{\overline{z}^N}\right)\hfill \\ & =\underset{N=0}{\overset{\mathrm{}}{}}\frac{(2\mathrm{})^{|N|}}{N!}\frac{^{|N|}(R_\alpha ^{}f)}{z^N}\frac{^{|N|}(R_\alpha ^{}g)}{\overline{z}^N}\hfill \\ & =R_\alpha ^{}f_{\mathrm{Wick}}^{\mathrm{}}R_\alpha ^{}g.\hfill \end{array}$$ The compatibility with the complex conjugation $`\overline{R_\alpha ^{}f}=R_\alpha ^{}\overline{f}`$ is obvious. ∎ ###### Remark 5.12 Of course we can also directly compare the seminorms for different values of $`\mathrm{}`$. Clearly, one has $$f_{0,0,R,S}^{0,\mathrm{}}f_{0,0,R,S}^{0,\mathrm{}^{}}$$ (5.19) for $`\mathrm{}\mathrm{}^{}`$ whence by induction $$f_{m,\mathrm{},R,S}^{0,\mathrm{}}f_{m,\mathrm{},R,S}^{0,\mathrm{}^{}}$$ (5.20) as well. For $`\mathrm{}\mathrm{}^{}`$ this gives the inclusion $$𝒜_{\mathrm{}^{}}𝒜_{\mathrm{}}.$$ (5.21) ## 6 The GNS construction and coherent states We shall now discuss the GNS construction corresponding to the positive $`\delta `$-functionals $$\delta _p:𝒜_{\mathrm{}},$$ (6.1) now in the convergent situation. ###### Proposition 6.1 Let $`p^n`$. Then the Gel’fand ideal of $`\delta _p`$ is given by $$𝒥_p=\{f𝒜_{\mathrm{}}|I:\frac{^{|I|}f}{\overline{z}^I}(p)=0\},$$ (6.2) and the GNS pre-Hilbert space $`𝔇_p=𝒜_{\mathrm{}}/𝒥_p`$ is a Fréchet space in the natural way, where the topology of $`𝔇_p`$ is determined by the seminorms $$[f]_{m,\mathrm{},R,S}^{p,\mathrm{}}=inf\left\{f+g_{m,\mathrm{},R,S}^{p,\mathrm{}}|g𝒥_p\right\}.$$ (6.3) ###### Proof. The statement (6.2) is obvious. Since $`\delta _p`$ is continuous, the Gel’fand ideal is a closed subspace of $`𝒜_{\mathrm{}}`$. Thus the quotient $`𝔇_p`$ is again a Fréchet space by general arguments, see e.g. \[29, Sect. 4.4, Prop. 1\]. ∎ Since the translation group acts by inner -automorphisms we can safely specialize to the case $`p=0`$ in a first step. Then we can describe the quotient $`𝔇_0`$ more explicitly: ###### Theorem 6.2 Let $`f,g𝒜_{\mathrm{}}`$. 1. The $`\overline{z}`$-Taylor expansion $`\mathrm{\Psi }:𝒜_{\mathrm{}}𝒜_{\mathrm{}}`$ defined by $$\mathrm{\Psi }:f\left(z\mathrm{\Psi }_f(z)=\underset{I=0}{\overset{\mathrm{}}{}}\frac{1}{I!}\frac{^{|I|}f}{\overline{z}^I}(0)\overline{z}^I\right)$$ (6.4) is a continuous projection with $$\mathrm{ker}\mathrm{\Psi }=𝒥_0.$$ (6.5) In fact, $$\mathrm{\Psi }_f_{m,\mathrm{},R,S}^{0,\mathrm{}}f_{m,\mathrm{},R,S}^{0,\mathrm{}},$$ (6.6) where equality holds if and only if $`f`$ is anti-holomorphic. 2. The quotient $`𝔇_0`$ is canonically isomorphic as a Fréchet space to the image $`𝔇=\mathrm{im}\mathrm{\Psi }`$ of $`\mathrm{\Psi }`$ via $$𝔇_0[f]\mathrm{\Psi }_f𝔇.$$ (6.7) In particular, $$[f]_{m,\mathrm{},R,S}^{0,\mathrm{}}=\mathrm{\Psi }_f_{m,\mathrm{},R,S}^{0,\mathrm{}}.$$ (6.8) 3. The space $`𝔇`$ is a dense subspace of the Bargmann-Fock Hilbert space $`_{\mathrm{BF}}`$. The map (6.7) is an isometry of pre-Hilbert spaces and the Fréchet topology of $`𝔇`$ is finer than the topology induced from the Hilbert space $`_{\mathrm{BF}}`$. 4. The GNS representation on $`𝔇_0`$ induces via (6.7) the Bargmann-Fock representation of $`𝒜_{\mathrm{}}`$ on $`𝔇`$, explicitly given by $$\pi (f)\mathrm{\Psi }_g=\underset{I=0}{\overset{\mathrm{}}{}}\frac{(2\mathrm{})^{|I|}}{I!}\left(\underset{J=0}{\overset{\mathrm{}}{}}\frac{1}{J!}\frac{^{|I|+|J|}f}{z^I\overline{z}^J}(0)\overline{z}^J\right)\frac{^{|I|}\mathrm{\Psi }_g}{\overline{z}^I},$$ (6.9) where both series converge in the topology of $`𝔇`$. 5. The bilinear map $`𝒜_{\mathrm{}}\times 𝔇(f,\mathrm{\Psi }_g)\pi (f)\mathrm{\Psi }_g𝔇`$ is continuous with respect to the Fréchet topologies of $`𝒜_{\mathrm{}}`$ and $`𝔇`$, respectively. ###### Proof. For the first part we have $`\mathrm{\Psi }_f(\overline{z})=\widehat{f}(0,\overline{z})`$ whence $`\mathrm{\Psi }_f`$ is indeed a well-defined anti-holomorphic function. Moreover, (6.6) is clear from our consideration in the proof of Theorem 3.9 since in the seminorm of $`\mathrm{\Psi }_f`$ simply less Taylor coefficients contribute compared to the corresponding seminorm of $`f`$. Thus $`\mathrm{\Psi }_f𝒜_{\mathrm{}}`$ and $`\mathrm{\Psi }`$ is continuous. From the explicit form of $`\mathrm{\Psi }`$ the equality (6.5) and $`\mathrm{\Psi }^2=\mathrm{\Psi }`$ are obvious. For the second part we first notice that (6.7) is well-defined and bijective since $`f\mathrm{\Psi }_f\mathrm{ker}\mathrm{\Psi }=𝒥_0`$ using the fact that $`\mathrm{\Psi }`$ is a projection. Moreover, (6.8) follows directly from (6.6) since obviously $`[f]_{m,\mathrm{},R,S}^{0,\mathrm{}}\mathrm{\Psi }_f_{m,\mathrm{},R,S}^{0,\mathrm{}}`$. Since the inverse of (6.7) is simply given by the well-defined and continuous map $`\mathrm{\Psi }_f[f]`$, we see that (6.7) is indeed an isomorphism of Fréchet spaces. For the third part we compute explicitly the GNS inner product on $`𝔇_0`$ $$\begin{array}{cc}\hfill [f],[g]_{𝔇_0}& =\delta _0(\overline{f}_{\mathrm{Wick}}^{\mathrm{}}g)\hfill \\ & =\underset{N=0}{\overset{\mathrm{}}{}}\frac{(2\mathrm{})^{|N|}}{N!}\frac{^{|N|}\overline{f}}{z^N}(0)\frac{^{|N|}g}{\overline{z}^N}(0)\hfill \\ & =\underset{N=0}{\overset{\mathrm{}}{}}\frac{(2\mathrm{})^{|N|}}{N!}\overline{\frac{^{|N|}\mathrm{\Psi }_f}{\overline{z}^N}(0)}\frac{^{|N|}\mathrm{\Psi }_g}{\overline{z}^N}(0)\hfill \\ & =\mathrm{\Psi }_f,\mathrm{\Psi }_g_{\mathrm{BF}},\hfill \end{array}$$ whence (6.7) is isometric. In particular $`𝔇_{\mathrm{BF}}`$ follows, as $$\mathrm{\Psi }_f,\mathrm{\Psi }_f_{\mathrm{BF}}=(\overline{f}_{\mathrm{Wick}}^{\mathrm{}}f)(0)=(f_{0,0,0,0}^{0,\mathrm{}})^2<\mathrm{}.$$ Since $`[\overline{z}]𝔇_{\mathrm{BF}}`$, the subspace $`𝔇`$ is dense in $`_{\mathrm{BF}}`$. Moreover, since $`f_{0,0,0,0}^{0,\mathrm{}}=\mathrm{\Psi }_f_{0,0,0,0}^{0,\mathrm{}}`$ the estimate ($``$) implies that the Fréchet topology of $`𝔇`$ is finer than the topology induced from $`_{\mathrm{BF}}`$. This shows the third part. For the fourth part recall that the GNS representation is defined by $`\varrho (f)[g]=[f_{\mathrm{Wick}}^{\mathrm{}}g]`$ which translates via (6.7) into $$\begin{array}{cc}\hfill \pi (f)\mathrm{\Psi }_g& =\mathrm{\Psi }_{f_{\mathrm{Wick}}^{\mathrm{}}g}\hfill \\ & =\mathrm{\Psi }_{_{N=0}^{\mathrm{}}\frac{(2\mathrm{})^{|N|}}{N!}\frac{^{|N|}f}{z^N}\frac{^{|N|}g}{\overline{z}^N}}\hfill \\ & \stackrel{(a)}{=}\underset{N=0}{\overset{\mathrm{}}{}}\frac{(2\mathrm{})^{|N|}}{N!}\mathrm{\Psi }_{\frac{^{|N|}f}{z^N}\frac{^{|N|}g}{\overline{z}^N}}\hfill \\ & \stackrel{(b)}{=}\underset{N=0}{\overset{\mathrm{}}{}}\frac{(2\mathrm{})^{|N|}}{N!}\mathrm{\Psi }_{\frac{^{|N|}f}{z^N}}\mathrm{\Psi }_{\frac{^{|N|}g}{\overline{z}^N}}\hfill \\ & =\underset{N=0}{\overset{\mathrm{}}{}}\frac{(2\mathrm{})^{|N|}}{N!}\left(\underset{M=0}{\overset{\mathrm{}}{}}\frac{1}{M!}\frac{^{|N|+|M|}f}{z^N\overline{z}^M}(0)\overline{z}^M\right)\frac{^{|N|}\mathrm{\Psi }_g}{\overline{z}^N},\hfill \end{array}$$ where $`(a)`$ holds since $`\mathrm{\Psi }`$ is continuous and $`(b)`$ holds since obviously $`\mathrm{\Psi }`$ is a homomorphism of the pointwise product. Finally, in the last step we have used that $`\mathrm{\Psi }`$ commutes with derivatives in $`\overline{z}`$-direction. This shows the fourth part and the last part follows immediately from $`\pi (f)\mathrm{\Psi }_g=\mathrm{\Psi }_{f_{\mathrm{Wick}}^{\mathrm{}}g}=\mathrm{\Psi }_{f_{\mathrm{Wick}}^{\mathrm{}}\mathrm{\Psi }_g}`$ and the continuity of $`\mathrm{\Psi }`$ and $`_{\mathrm{Wick}}^{\mathrm{}}`$. ∎ ###### Corollary 6.3 The Bargmann-Fock representation of $`𝒜_{\mathrm{}}`$ is injective. The following corollary is remarkable in so far as closed subspaces of Fréchet spaces usually do not have complementary closed subspaces: ###### Corollary 6.4 The algebra $`𝒜_{\mathrm{}}`$ decomposes into two complementary closed subspaces $`𝒜_{\mathrm{}}=𝔇𝒥_0`$. ###### Remark 6.5 Since the Bargmann-Fock representation is injective and since $$\pi (z^i)=2\mathrm{}\frac{}{\overline{z}^i}=a_i$$ (6.9) $$\pi (\overline{z}^i)=\overline{z}^i=a_i^{}$$ (6.10) are the annihilation and creation operators we find another interpretation of the algebra $`𝒜_{\mathrm{}}`$: it is a (rather large) completion of the polynomials in the creation and annihilation operators in a certain Fréchet topology. In particular, this completion contains the usual unitary generators of the Weyl algebra, i.e. the exponential functions of $`a_i`$ and $`a_i^{}`$. The are given by $`\pi (𝖾_{\overline{\alpha },\beta })`$ for suitable $`\alpha ,\beta ^n`$. ###### Remark 6.6 Since the Bargmann-Fock representation is a -representation of the -algebra $`𝒜_{\mathrm{}}`$ by (in general) unbounded operators with common domain $`𝔇`$, one can investigate the resulting $`O^{}`$-algebra of unbounded operators by techniques as developed in e.g. . In particular, it would be interesting to find more concrete characterizations of the Fréchet topologies of $`𝔇`$ and $`𝒜_{\mathrm{}}`$. We now discuss the action of the translation group $`^n`$ and its central extension $`𝖧_n`$. The following statement is obvious, the representation itself being well-known: ###### Lemma 6.7 The map $$𝖧_n(w,c)U_{(w,c)}=\pi (u_{(w,c)})\mathrm{U}(_{\mathrm{BF}})$$ (6.11) is a strongly continuous unitary representation of the Heisenberg group. Explicitly, $$(U_{(w,c)}\psi )(\overline{z})=\mathrm{e}^{\frac{\mathrm{i}}{2\mathrm{}}c\frac{\overline{w}w}{2\mathrm{}}}\mathrm{e}^{\frac{w\overline{z}}{2\mathrm{}}}\psi (\overline{z}+\overline{w})$$ (6.12) for $`\psi _{\mathrm{BF}}`$. It factors to a projective representation $`U_w=U_{(w,0)}`$ of the translation group $`^n`$. ###### Proof. Since $`\pi `$ is a -representation it follows that $`\pi (u_{(w,c)})`$ is a unitary operator defined on the dense domain $`𝔇`$. Thus it extends to a unitary operator on $`_{\mathrm{BF}}`$. The group representation property is obvious from (5.10). The explicit formula is a simple consequence of (6.9). The fact that (6.11) is strongly continuous is well-known but can also be shown within our approach directly: let $`\varphi ,\psi 𝔇`$ then $`g(w,c)=\psi ,U_{(w,c)}\varphi _{\mathrm{BF}}`$ is real-analytic since $`,_{\mathrm{BF}}`$ and $`\pi `$ are continuous with respect to the Fréchet topology and $`(w,c)u_{(w,c)}`$ is real-analytic according to Theorem 5.8. Since $`g(0,0)=\psi ,\varphi _{\mathrm{BF}}`$ we see that on the dense domain $`𝔇`$ the representation (6.11) is weakly continuous at the identity. But this implies that it is strongly continuous on the whole group $`𝖧_n`$ and on the whole Hilbert space $`_{\mathrm{BF}}`$. Since the contribution of $`c`$ is only an overall phase, the representation clearly factors to a projective representation of $`^n`$. ∎ Since we have a group action of $`𝖧_n`$ we can formulate now the following covariance property of the Bargmann-Fock representation which is an obvious consequence of the fact that the -automorphisms are *inner*. ###### Theorem 6.8 The Bargmann-Fock representation is $`𝖧_n`$-covariant with respect to the action by -automorphisms on $`𝒜_{\mathrm{}}`$ and the action by unitaries on $`_{\mathrm{BF}}`$, i.e. we have $$\pi (\mathrm{Ad}_{\mathrm{Wick}}(u_{(w,c)})f)=U_{(w,c)}\pi (f)U_{(w,c)}^{}=U_w\pi (f)U_w^{}$$ (6.13) for all $`(w,c)𝖧_n`$ and $`f𝒜_{\mathrm{}}`$. Since the GNS representation is cyclic with cyclic vector $`\mathrm{\Psi }_1=1`$ we obtain *coherent states* with respect to the representation of $`𝖧_n`$. We define the *coherent state vector* $`\psi _{(w,c)}_{\mathrm{BF}}`$ by $$\psi _{(w,c)}=U_{(w,c)}^1\psi _1$$ (6.14) explicitly given by $$\psi _{(w,c)}(\overline{z})=\mathrm{e}^{\frac{\mathrm{i}}{2\mathrm{}}c+\frac{\overline{w}w}{4\mathrm{}}}\mathrm{e}^{\frac{w\overline{z}}{2\mathrm{}}}.$$ (6.15) From the covariance property (6.13) we immediately have the following characterization of the $`\delta `$-functionals at arbitrary points in $`^n`$: ###### Corollary 6.9 The coherent state vectors give the $`\delta _p`$-functionals as expectation value functionals, i.e. we have $$\delta _w(f)=\psi _{(w,c)},\pi (f)\psi _{(w,c)}_{\mathrm{BF}}$$ (6.16) for all $`f𝒜_{\mathrm{}}`$ and $`(w,c)𝖧_n`$. In particular, the group action of $`𝖧_n`$ on the coherent state vectors factors through to a group action of the translation group $`^n`$ on the coherent states $`\delta _w:𝒜_{\mathrm{}}`$. ###### Remark 6.10 This corollary gives finally the justification to view the $`\delta `$-functionals of the Wick star product algebra as coherent states with respect to the translation group. Though the explicit formula (6.15) is folklore (it is just the Bergmann kernel) we would like to emphasize that in our approach the coherent states emerge out of properties of the observable algebra instead of more conventional approaches based on group actions on the state vectors in some Hilbert space, see e.g. for more references. In this sense our approach supports the idea that the observable algebra is the more fundamental object in both, quantum and classical mechanics. ###### Remark 6.11 We also note that the statement of Theorem 6.8 as well as the Corollary 6.9 are *not* possible for the formal Bargmann-Fock representation. Again, this was one of our main motivations to consider a suitable convergence scheme for the Wick star product. The result of Theorem 6.8 and Corollary 6.9 suggest the following general definition of coherent states with respect to some symmetry based on the *observable algebra*: ###### Definition 6.12 Let $`𝒜`$ be a -algebra with unit $`\mathrm{𝟙}`$ and let $`G`$ be a group acting on $`𝒜`$ by -automorphisms $`\mathrm{\Phi }_g:𝒜𝒜`$. Let $`\omega `$ be a state of $`𝒜`$ such that the GNS representation is $`G`$-covariant, i.e. there exists a unitary (or more general: projectively unitary) representation $`U`$ of $`G`$ on the GNS pre-Hilbert space $`_\omega `$. Then the states $`\omega _g`$ with $$\omega _g(a)=(\omega \mathrm{\Phi }_g)(a)=\psi _g,\pi (a)\psi _g,$$ (6.17) where $`\psi _g=U_g^{}\psi _\mathrm{𝟙}_\omega `$ are called coherent with respect to $`G`$. Clearly, in case of a projective representation only the coherent states $`\omega _g`$ are well-defined, while for a unitary representation also the coherent state vectors $`\psi _g`$ are well-defined. With the Heisenberg group acting on $`𝒜_{\mathrm{}}`$ and the $`\delta `$-functional we are in this situation: if the action is realized by *inner* -automorphisms then the representation is always covariant for a (in general only projective) representation on the GNS pre-Hilbert space. Note also, that for an *invariant* state $`\omega `$ the GNS representation is trivially covariant. In this case $`\psi _g=\psi _\mathrm{𝟙}`$ coincides with the vacuum vector for all $`gG`$. Thus the interesting coherent states arise from non-invariant vacua such that the GNS representation is nevertheless covariant. For a survey on covariant -representation theory for -algebras over ordered rings we refer to . Let us now come to a further property of the subspace $`𝔇_{\mathrm{BF}}`$. We have already seen that the action of $`𝖧_n`$ leaves $`𝔇`$ invariant. ###### Theorem 6.13 The vectors in $`𝔇`$ are analytic with respect to the unitary representation $`U`$ of $`𝖧_n`$. ###### Proof. Let $`\psi 𝔇`$ be fixed then we have to show that $`𝖧_n(w,c)U_{(w,c)}\psi `$ is analytic with respect to the topology of $`_{\mathrm{BF}}`$. But this is simple since $`U_{(w,c)}\psi =\pi (u_{(w,c)})\psi `$ and $`(w,c)u_{(w,c)}`$ is analytic in the topology of $`𝒜_{\mathrm{}}`$. Moreover, the *bilinear* map $`(f,\psi )\pi (f)\psi `$ is continuous in the topologies of $`𝒜_{\mathrm{}}`$ and $`𝔇`$. Thus the map $`(w,c)U_{(w,c)}\psi `$ is analytic with respect to the topology of $`𝔇`$. Since by Theorem 6.2 this topology is finer than the one of $`_{\mathrm{BF}}`$, the proof is complete. ∎ Thus it would be interesting to know whether $`𝔇`$ coincides with the space of all analytic vectors. A positive answer would help to understand the (still rather complicated) Fréchet topology of $`𝔇`$ and hence the one of $`𝒜_{\mathrm{}}`$. Using the convergence of the star exponential (5.15) we even can specify the analyticity of the vectors in $`𝔇`$ further. Since $`\pi `$ is continuous with respect to the topologies of $`𝒜_{\mathrm{}}`$ and $`𝔇`$, we have $$U_{(w,c)}\psi =\pi (u_{(w,c)})\psi =\pi \left(\mathrm{Exp}(J_{(w,c)})\right)\psi =\underset{r=0}{\overset{\mathrm{}}{}}\frac{1}{r!}\pi \left(J_{(w,c)}\right)^r\psi $$ (6.18) with respect to the topology of $`𝔇`$. Again, since the topology of $`𝔇`$ is finer than the one of $`_{\mathrm{BF}}`$, the series converges unconditionally also in the Hilbert space sense. As usual, the contribution of $`c`$ is not essential. ###### Corollary 6.14 Let $`\psi 𝔇`$ and $`w^n`$. Then the series $$U_w\psi =\underset{r=0}{\overset{\mathrm{}}{}}\frac{1}{r!}\left(\frac{1}{2\mathrm{}}\right)^r\left(\underset{i=1}{\overset{n}{}}\overline{w}^ia_iw^ia_i^{}\right)^r\psi $$ (6.19) converges unconditionally in the Hilbert space topology. Of course we can also rewrite this in terms of the position and momentum operators $$Q_k=\frac{1}{2}(a_k+a_k^{})=\frac{1}{2}\pi (z^k+\overline{z}^k)$$ (6.20) and $$P_k=\frac{1}{2\mathrm{i}}(a_ka_k^{})=\frac{1}{2\mathrm{i}}\pi (z^k\overline{z}^k),$$ (6.21) defined as unbounded symmetric operators on $`𝔇`$. Then (6.18) shows that for $`\psi 𝔇`$ we have the unconditionally convergent series $$\mathrm{e}^{\frac{\mathrm{i}\stackrel{}{p}\stackrel{}{Q}}{\mathrm{}}}\psi =\underset{r=0}{\overset{\mathrm{}}{}}\frac{1}{r!}\left(\frac{\mathrm{i}\stackrel{}{p}\stackrel{}{Q}}{\mathrm{}}\right)^r\psi $$ (6.22) and $$\mathrm{e}^{\frac{\mathrm{i}\stackrel{}{q}\stackrel{}{P}}{\mathrm{}}}\psi =\underset{r=0}{\overset{\mathrm{}}{}}\frac{1}{r!}\left(\frac{\mathrm{i}\stackrel{}{q}\stackrel{}{P}}{\mathrm{}}\right)^r\psi $$ (6.23) in the Hilbert space topology, for all $`\stackrel{}{q},\stackrel{}{p}^n`$ substituting $`w`$ suitably. Of course, this also follows by ‘Hilbert space techniques’ from the strong continuity of the representation $`U`$ and Theorem 6.13. Note however, that the above argument using the convergence of the star exponential is independent.
warning/0506/math0506466.html
ar5iv
text
# + + = (Formulas of Brion, Lawrence, and Varchenko on rational generating functions for cones) ## Proofs Brion’s original proof of his formula used the Lefschetz–Riemann–Roch theorem in equivariant $`K`$-theory applied to a singular toric variety. Fortunately for us, the remarkable formulas of Brion and of Lawrence–Varchenko now have easy proofs, based on counting. Let us first consider an example based on the cone $`𝒦=_0(0,1)+_0(2,1)`$. The open circles in the picture on the left in Figure 1 represent the semigroup $`(0,1)+(2,1)`$, which is a proper subsemigroup of the integer points $`𝒦^2`$ in $`𝒦`$. The picture on the right shows how translates of the fundamental half-open parallelepipied $`𝒫`$ by this subsemigroup cover $`𝒦`$. This gives the formula $$\sigma _𝒦(x)=\sigma _𝒫(x)\underset{m,n0}{}x^m(x^2y)^n=\frac{1+xy}{(1x)(1x^2y)},$$ as the fundamental parallelepiped $`𝒫`$ contains two integer points, the origin and the point $`(1,1)`$. A simple rational cone in $`^d`$ has the form $$𝒦:=\{𝐯+\underset{i=1}{\overset{d}{}}\lambda _i𝐰_i\lambda _i_0\}=𝐯+\underset{i=1}{\overset{d}{}}_0𝐰_i,$$ where $`𝐰_1,\mathrm{},𝐰_d^d`$ are linearly independent. This cone is tiled by the $`(𝐰_1+\mathrm{}+𝐰_d)`$-translates of the half-open parallelepiped $$𝒫:=\{𝐯+\underset{i=1}{\overset{d}{}}\lambda _i𝐰_i0\lambda _i<1\}.$$ The generating function for $`𝒫`$ is the polynomial $$\sigma _𝒫(x)=\underset{𝐦𝒫^d}{}x^𝐦,$$ and so the generating function for $`𝒦`$ is $$\sigma _𝒦(x)=\underset{\alpha 𝐰_1+\mathrm{}+𝐰_d}{}x^\alpha \sigma _𝒫(x)=\frac{\sigma _𝒫(x)}{(1x^{𝐰_1})\mathrm{}(1x^{𝐰_d})},$$ which is a rational function. This formula and its proof do not require that the apex $`𝐯`$ be rational, but only that the generators $`𝐰_i`$ of the cone be linearly independent vectors in $`^d`$. A rational cone $`𝒦`$ with apex $`𝐯`$ and generators $`𝐰_1,\mathrm{},𝐰_n^d`$ has the form $$𝒦=𝐯+_0𝐰_1+\mathrm{}+_0𝐰_n.$$ If there is a vector $`\xi ^d`$ with $`\xi 𝐰_i>0`$ for $`i=1,\mathrm{},n`$, then $`𝒦`$ is strictly convex. A fundamental result on convexity \[2, Lemma VIII.2.3\] is that $`𝒦`$ may be decomposed into simple cones $`𝒦_1,\mathrm{},𝒦_l`$ having pairwise disjoint interiors, each with apex $`𝐯`$ and generated by $`d`$ of the generators $`𝐰_1,\mathrm{},𝐰_n`$ of $`𝒦`$. We would like to add the generating functions for each cone $`𝒦_i`$ to obtain the generating function for $`𝒦`$. However, some of the cones may have lattice points in common, and some device is needed to treat the subsequent overcounting. An elegant way to do this is to avoid the overcounting altogether by translating all the cones . We explain this. There exists a short vector $`𝐬^d`$ such that (4) $$𝒦^d=(𝐬+𝒦)^d,$$ and no facet of any cone $`𝐬+𝒦_1,\mathrm{},𝐬+𝒦_l`$ contains any integer points. This gives the disjoint irrational decomposition $$𝒦^d=(𝐬+𝒦_1)^d\mathrm{}(𝐬+𝒦_l)^d,$$ and so (5) $$\sigma _𝒦(x)=\underset{𝐦𝒦^d}{}x^𝐦=\underset{i=1}{\overset{l}{}}\sigma _{𝐬+𝒦_i}(x)$$ is a rational function. For example, suppose that $`𝒦`$ is the cone in $`^3`$ with apex the origin and generators $$𝐰_1=(1,0,1),𝐰_2=(0,1,1),𝐰_3=(0,1,1),\text{and}𝐰_4=(1,0,1).$$ If we let $`𝒦_1`$ be the simple cone with generators $`𝐰_1,𝐰_2,𝐰_3`$ and $`𝒦_2`$ be the simple cone with generators $`𝐰_2,𝐰_3,𝐰_4`$, then $`𝒦_1`$ and $`𝒦_2`$ decompose $`𝒦`$ into simple cones. If $`𝐬=(\frac{1}{8},0,\frac{1}{3})`$, then (4) holds, and no facet of $`𝐬+𝒦_1`$ or of $`𝐬+𝒦_2`$ contains any integer points. We display these cones, together with their integer points having $`z`$-coordinate 0, 1, or 2. The cone $`𝐬+𝒦_1`$ contains the 5 magenta points shown with positive first coordinate, while $`𝐬+𝒦_2`$ contains the other displayed points. Their integer generating functions are $`\sigma _{𝐬+𝒦_1}(x)`$ $`=`$ $`{\displaystyle \frac{x+xz}{(1yz)(1y^1z)(1xz)}},`$ $`\sigma _{𝐬+𝒦_2}(x)`$ $`=`$ $`{\displaystyle \frac{1+z}{(1yz)(1y^1z)(1x^1z)}},\text{and}`$ $`\sigma _𝒦(x)`$ $`=`$ $`{\displaystyle \frac{(1+x)(1z^2)}{(1yz)(1y^1z)(1xz)(1x^1z)}}.`$ Then $`\sigma _{𝐬+𝒦_1}(x)+\sigma _{𝐬+𝒦_2}(x)=\sigma _𝒦(x)`$, as $$(x+xz)(1x^1z)+(1+z)(1xz)=1+xz^2xz^2=(1+x)(1z^2).$$ While the cones that appear in the Lawrence–Varchenko formula are all simple, and those in Brion’s formula are strictly convex, we use yet more general cones in their proof. A rational (closed) halfspace is the convex subset of $`^d`$ defined by $$\{x^d𝐰xb\},$$ where $`𝐰^d`$ and $`b`$. Its boundary is the rational hyperplane $`\{x^d𝐰x=b\}`$. A (closed) cone $`𝒦`$ is the interection of finitely many closed halfspaces whose boundary hyperplanes have some point in common. We assume this intersection is irredundant. The apex of $`𝒦`$ is the intersection of these boundary hyperplanes, which is an affine subspace. The generating function for the integer points in $`𝒦`$ is the formal Laurent series (6) $$S_𝒦:=\underset{𝐦𝒦}{}x^𝐦.$$ This formal series makes sense as a rational function only if $`𝒦`$ is strictly convex, that is, if its apex is a single point. Otherwise, the apex is a rational affine subspace $`L`$, and the cone $`𝒦`$ is stable under translation by any integer vector $`𝐰`$ that is parallel to $`L`$. If $`𝐦𝒦^d`$, then the series $`S_𝒦`$ contains the series $$x^𝐦\underset{n}{}x^{n𝐰}$$ as a subsum. As this converges only for $`x=0`$, the series $`S_𝒦`$ converges only for $`x=0`$. We relate these formal Laurent series to rational functions. The product of a formal series and a polynomial is another formal series. Thus the additive group $`[[x_1^{\pm 1},\mathrm{},x_d^{\pm 1}]]`$ of formal Laurent series is a module over the ring $`[x_1^{\pm 1},\mathrm{},x_d^{\pm 1}]`$ of Laurent polynomials. The space $`\mathrm{PL}`$ of polyhedral Laurent series is the $`[x_1^{\pm 1},\mathrm{},x_d^{\pm 1}]`$-submodule of $`[[x_1^{\pm 1},\mathrm{},x_d^{\pm 1}]]`$ generated by the set of formal series $$\{S_𝒦𝒦\text{ is a simple rational cone}\}.$$ Since any rational cone may be triangulated by simple cones, $`\mathrm{PL}`$ contains the integer generating series of all rational cones. Let $`(x_1,\mathrm{},x_d)`$ be the field of rational functions on $`^d`$, which is the quotient field of $`[x_1^{\pm 1},\mathrm{},x_d^{\pm 1}]`$. According to Ishida , the proof of the following theorem is due to Brion. ###### Theorem 7. There is a unique homomorphism of $`[x_1^{\pm 1},\mathrm{},x_d^{\pm 1}]`$-modules $$\phi :\mathrm{PL}(x_1,\mathrm{},x_d),$$ such that $`\phi (S_𝒦)=\sigma _𝒦`$ for every simple cone $`𝒦`$ in $`^d`$. Proof. Given a simple rational cone $`𝒦=v+𝐰_1,\mathrm{},𝐰_d`$ with fundamental parallelepiped $`𝒫`$, we have $$\underset{i=1}{\overset{d}{}}(1x^{𝐰_i})S_𝒦=\sigma _𝒫(x).$$ Hence, for each $`S\mathrm{PL}`$, there is a nonzero Laurent polynomial $`g[x_1^{\pm 1},\mathrm{},x_d^{\pm 1}]`$ such that $`gS=f[x_1^{\pm 1},\mathrm{},x_d^{\pm 1}]`$. If we define $`\phi (S):=f/g(x_1,\mathrm{},x_d)`$, then $`\phi (S)`$ is independent of the choice of $`g`$. This defines the required homomorphism. $`\mathrm{}`$ The map $`\phi `$ takes care of the nonconvergence of the generating series $`S_𝒦`$ when $`𝒦`$ is not strictly convex. ###### Lemma 8. If a rational polyhedral cone $`𝒦`$ is not strictly convex, then $`\phi (S_𝒦)=0`$. Proof. Let $`𝒦`$ be a rational polyhedral cone that is not strictly convex. Then there is a nonzero vector $`𝐰^d`$ such that $`𝐰+𝒦=𝒦`$, and so $`x^𝐰S_𝒦=S_𝒦`$. Thus $`x^𝐰\phi (S_𝒦)=\phi (S_𝒦)`$. Since $`1x^𝐰`$ is not a zero-divisor in $`(x_1,\mathrm{},x_d)`$, we conclude that $`\phi (S_𝒦)=0`$. $`\mathrm{}`$ We now establish Brion’s Formula, first for a simplex, and then use irrational decomposition for the general case. (A $`d`$-dimensional simplex is the intersection of $`d+1`$ halfspaces, one for each facet.) For a face $`F`$ of the simplex $`𝒫`$, let $`𝒦_F`$ be the tangent cone to $`F`$, which is the intersection of the halfspaces corresponding to the $`ddim(F)`$ facets containing $`F`$. Let $`\mathrm{}`$ be the empty face of $`𝒫`$, which has dimension $`1`$. Its tangent cone is $`𝒫`$. ###### Theorem 9. If $`P`$ is a simplex, then (10) $$0=\underset{F}{}(1)^{dim(F)}S_{𝒦_F},$$ the sum over all faces of $`P`$. Proof. Consider the coefficient of $`x^𝐦`$ for some $`𝐦^d`$ in the sum on the right. Then $`𝐦`$ lies in the tangent cone $`𝒦_F`$ to a unique face $`F`$ of minimal dimension, as $`P`$ is a simplex. The coefficient of $`x^𝐦`$ in the sum becomes $$\underset{GF}{}(1)^{dim(G)}.$$ But this vanishes, as every interval in the face poset of $`P`$ is a Boolean lattice. $`\mathrm{}`$ Now we apply the evaluation map $`\phi `$ of Theorem 7 to the formula (10). Lemma 8 implies that $`\phi (S_{𝒦_F})=0`$ except when $`F=\mathrm{}`$ or $`F`$ is a vertex, and then $`\phi (S_{𝒦_F})=\sigma _{𝒦_F}(x)`$. This gives $$0=\sigma _𝒫(x)+\underset{𝐯\text{ a vertex of }𝒫}{}\sigma _{𝒦_𝐯}(x),$$ which is Brion’s Formula for simplices. Just as for rational cones, every polytope $`𝒫`$ may be decomposed into simplices $`𝒫_1,\mathrm{},𝒫_l`$ having pairwise disjoint interiors, using only the vertices of $`𝒫`$. $$𝒫=𝒫_1\mathrm{}𝒫_l.$$ Then there exists a small real number $`ϵ>0`$ and a short vector $`𝐬`$ such that if we set $$𝒫^{}:=𝐬+(1+ϵ)𝒫\text{and}𝒫_i^{}:=𝐬+(1+ϵ)𝒫_i\text{for}i=1,\mathrm{},l,$$ then $`𝒫^{}^d=𝒫^d`$, and no hyperplane supporting any facet of any simplex $`𝒫_i^{}`$ meets $`^d`$. If we write $`𝒦(𝒬)_𝐰`$ for the tangent cone to a polytope $`𝒬`$ at a vertex $`𝐰`$, then for $`𝐯`$ a vertex of $`𝒫`$ with $`𝐯^{}=(1+ϵ)𝐯+𝐬`$ the coresponding vertex of $`𝒫^{}`$, we have $`𝒦(𝒫^{})_𝐯^{}^d=𝒦(𝒫)_𝐯^d`$ and so this is an irrational decomposition. Then $`{\displaystyle \underset{𝐯\text{ a vertex of }𝒫}{}}\sigma _{𝒦(𝒫)_𝐯}(x)`$ $`=`$ $`{\displaystyle \underset{𝐯\text{ a vertex of }𝒫^{}}{}}\sigma _{𝒦(𝒫^{})_𝐯}(x)`$ $`=`$ $`{\displaystyle \underset{i=1}{\overset{l}{}}}{\displaystyle \underset{𝐯\text{ a vertex of }𝒫_i^{}}{}}\sigma _{𝒦(𝒫_i^{})_𝐯}(x)`$ $`=`$ $`{\displaystyle \underset{i=1}{\overset{l}{}}}\sigma _{𝒫_i}(x)=\sigma _𝒫^{}(x)=\sigma _𝒫(x).`$ The second equality holds because the vertex cones $`𝒦(𝒫_i^{})_𝐯`$ form an irrational decomposition of the vertex cone $`𝒦(𝒫^{})_𝐯`$, and because the same is true for the polytopes. This completes our proof of Brion’s Formula. Consider the quadrilateral $`𝒬`$, which may be triangulated by adding an edge between the vertices $`(2,0)`$ and $`(0,2)`$. Let $`ϵ=\frac{1}{4}`$ and $`𝐬=(\frac{1}{2},\frac{1}{4})`$. Then $`(1+ϵ)𝒬+𝐬`$ has vertices $$(\frac{1}{2},\frac{1}{4}),(2,\frac{1}{4}),(\frac{1}{2},2+\frac{1}{4}),(4+\frac{1}{2},2+\frac{1}{4}).$$ We display the resulting irrational decomposition. We use the map $`\phi `$ to deduce a very general form of the Lawrence–Varchenko formula. Let $`𝒫`$ be a simple polytope, and for each vertex $`𝐯`$ of $`𝒫`$ choose a vector $`\xi _𝐯`$ that is not perpendicular to any edge direction at $`𝐯`$. Form the cone $`𝒦_{\xi _𝐯,𝐯}`$ as before. Then we have (11) $$\sigma _𝒫(x)=\underset{𝐯\text{ a vertex of }𝒫}{}(1)^{|E_𝐯^{}(\xi _𝐯)|}\sigma _{𝒦_{\xi _𝐯,𝐯}}(x).$$ Brion’s formula is the special case when each vector $`\xi _𝐯`$ points into the interior of the polytope. We establish (11) by showing that the sum on the right does not change when any of the vectors $`\xi _𝐯`$ are rotated. Pick a vertex $`𝐯`$ and vectors $`\xi ,\xi ^{}`$ that are not perpendicular to any edge direction at $`𝐯`$ such that $`\xi 𝐰`$ and $`\xi 𝐰^{}`$ have the same sign for all except one edge direction $`𝐦`$ at $`𝐯`$. Then $`𝒦_{\xi ,𝐯}`$ and $`𝒦_{\xi ^{},𝐯}`$ are disjoint and their union is the (possibly) half-open cone $`𝒦`$ generated by the edge directions $`𝐰`$ at $`𝐯`$ such that $`\xi 𝐰`$ and $`\xi ^{}𝐰`$ have the same sign, but with apex the affine line $`𝐯+𝐦`$. Thus we have the identity of rational formal series $$S_{𝒦_{\xi ,𝐯}}S_𝒦=S_{𝒦_{\xi ^{},𝐯}}.$$ Applying the evaluation map $`\phi `$ gives $$\sigma _{𝒦_{\xi ,𝐯}}(x)=\sigma _{𝒦_{\xi ^{},𝐯}}(x),$$ which proves the claim, and the generalized Lawrence–Varchenko formula (11). ## Valuations Valuations provide a conceptual approach to these ideas. Once the theory is set up, both Brion’s Formula and the Lawrence–Varchenko Formula are easy corollaries of duality being a valuation. We are indebted to Sasha Barvinok who pointed out this correspondence to the second author during a coffee break at the 2005 Park City Mathematical Institute. Let us explain. Consider the vector space of all functions $`^d`$. Let $`𝒱`$ be the subspace that is generated by indicator functions of polyhedra: $$[𝒫]:x\{\begin{array}{cc}1\hfill & \text{ if }x𝒫,\hfill \\ 0\hfill & \text{ if }x𝒫.\hfill \end{array}$$ We add these functions point-wise. For example, if $`d=1`$, and $`𝒫=[0,2]`$, $`𝒬=[1,3]`$, then $`[𝒫]+[𝒬]`$ takes the value $`1`$ along $`[0,1)`$ and $`(2,3]`$, the value $`2`$ along $`[1,2]`$, and vanishes everywhere else. + = Already this simple example shows that our generators do not form a basis: they are linearly dependent. For $`𝒫^{}=[0,3]`$ and $`𝒬^{}=[1,2]`$, we get the same sum. + = But this is the only thing that can happen. ###### Theorem 12 (). The linear space of relations among the indicator functions $`[𝒫]`$ of convex polyhedra is generated by the relations $`[𝒫]+[𝒬]=[𝒫𝒬]+[𝒫𝒬]`$ where $`𝒫`$ and $`𝒬`$ run over polyhedra for which $`𝒫𝒬`$ is convex. A valuation is a linear map $`\nu :𝒱V`$, where $`V`$ is some vector space. Some standard examples are | $`V`$ | $`\nu (𝒫)`$ | | --- | --- | | $`^d`$ | $`\mathrm{vol}(𝒫)`$ | | $`\mathrm{PL}`$ | $`S_𝒫(x)`$ | | $`(x_1,\mathrm{},x_d)`$ | $`\sigma _𝒫(x)`$ | | $`^d`$ | 1 | . That $`\sigma _𝒫(x)`$ is a valuation is a deep result of Khovanskii-Pukhlikov and of Lawrence . The last example is called the Euler characteristic. This valuation is surprisingly useful. For example, it can be used to prove Theorem 13 below. The most interesting valuation for us comes from the polar construction. The polar $`𝒫^{}`$ of a polyhedron $`𝒫`$ is the polyhedron given by $$𝒫^{}:=\{xx,y1\text{ for all }y𝒫\}.$$ It is instructive to work through some examples. 1. | | | | | --- | --- | --- | | The polar of the square | … | is the diamond. | 2. | | | | | --- | --- | --- | | The polar of a cone $`𝒦`$ | … | is the cone $`𝒦^{}:=\{xx,y0\text{ for all }y𝒦\}`$ . | 3. Suppose that $`𝒫`$ is a polytope whose interior contains the orign and $``$ is a face of $`𝒫`$. Then the polar of the tangent cone $`𝒦_{}`$ is the convex hull of the origin together with the dual face $`^{}:=\{x𝒫^{}x,y=1\}`$, which is a pyramid over $`^{}`$. For this last remark, note that if $`x^{}`$ and $`y𝒦_{}`$, then $`x,y^{},=1`$. Conversely, if $`x𝒦_{}^{}`$, then $`x,\text{.}`$ is maximized over $`𝒦_{}`$ at $``$ by example (2), and it is at most $`1`$ there. In these examples, the polar of the polar is the original polyhedron. This happens if and only if the original polyhedron contains the origin. 1. The polar of the interval $`[1,2]`$ is the interval $`[0,1/2]`$, but the polar of $`[0,1/2]`$ is $`[0,2]`$. Now, we come to the main theorem of this section. ###### Theorem 13 (Lawrence ). The assignment $`[𝒫][𝒫^{}]`$ defines a valuation. This innocent-looking result has powerful consequences. Suppose that $`𝒫`$ is a polytope whose interior contains the orign. Then we can cover $`𝒫^{}`$ by pyramids $`\mathrm{conv}(0,^{})`$ over the codimension-one faces $`^{}`$ of $`𝒫^{}`$. The indicator functions of $`𝒫`$ and the cover differ by indicator functions of pyramids of smaller dimension. (14) $$[𝒫^{}]=\underset{^{}}{}[\mathrm{conv}(0,^{})]\pm \text{ lower dimensional pyramids}.$$ The Euler–Poincaré formula for general polytopes organizes this inclusion-exclusion, giving the exact expression $$[𝒫^{}]=(1)^{\mathrm{codim}^{}+1}[\mathrm{conv}(0,^{})].$$ We illustrate this when $`𝒫`$ is the square. $`=`$ = $`+`$$`+`$$`+`$ $``$ $``$$``$$``$$`+`$. If we apply polarity to (14), we get the Brianchon–Gram Theorem . (15) $$[𝒫]=\underset{v\text{ vertex}}{}[𝒦_v]\pm \text{ tangent cones of faces of positive dimension}.$$ This is essentially the indicator function version of Theorem 9, but for general polytopes. If we now apply the valuation $`\sigma `$, and recall that $`\sigma `$ evaluates to zero on cones that are not strictly convex, we obtain Brion’s Formula. Next, suppose that we are given a generic direction vector $`\xi `$. On a face $``$ of $`𝒫`$, the dot product with $`\xi `$ achieves its maximum at a vertex $`v_\xi ()`$. For a vertex $`v`$ of $`𝒫`$, we set $$_\xi ^{}(v):=\underset{:v_\xi ()=v}{}\mathrm{relint}^{}.$$ (The relative interior, $`\mathrm{relint}(𝒫)`$, of a polyhedron $`𝒫`$ is the topological interior when considered as a subspace of its affine hull.) In words, we attach the relative interior of a low-dimensional pyramid $`\mathrm{conv}(0,^{})`$ to the full-dimensional pyramid $`\mathrm{conv}(0,v^{})`$ which we see when we look in the $`\xi `$-direction from $`\mathrm{conv}(0,^{})`$. In this way, we obtain an honest decomposition (16) $$[𝒫^{}]=\underset{v}{}[\mathrm{conv}(0,_\xi ^{}(v))].$$ For the polar of the square, this is $`=`$ $`=`$ $`+`$ $`+`$ $`+`$ . To compute the polar of the half-open polyhedron $`\mathrm{conv}(0,_\xi ^{}(v))`$, we have to write its indicator function $`[\mathrm{conv}(0,_\xi ^{}(v))]`$ as a linear combination of indicator functions of (closed) polyhedra. If $`𝒫`$ is a simple polytope, then all the dual faces $`^{}`$ are simplices. It turns out that the polar of $`\mathrm{conv}(0,_\xi ^{}(v))`$ is precisely the forward tangent cone $`𝒦_{\xi ,v}`$ at the vertex $`v`$. So the Lawrence–Varchenko formula is just the polar of (16). This gives a fairly general principle to construct Brion-type formulas: Choose a decomposition of (the indicator function of) $`𝒫^{}`$, and then polarize. We invite the reader to set up their own equations this way. ## An Application Brion’s Formula shows that certain data of a *polytope*—the list of its integer points encoded in a generating function—can be reduced to *cones*. We have already seen how to construct the generating function $`\sigma _𝒦(x)`$ for a simple cone $`𝒦`$. General cones can be composed from simple ones via triangulation and either irrational decomposition or inclusion-exclusion. Given a rational polytope $`𝒫`$, Brion’s Formula allows us to write the possibly huge polynomial $`\sigma _𝒫(x)`$ as a sum of rational functions, which stem from (triangulations of) the vertex cones. A priori it is not clear that this rational-function representation of $`\sigma _𝒫(x)`$ is any shorter than the original polynomial. That this is indeed possible is due to the *signed decomposition* theorem of Barvinok . To state Barvinok’s Theorem, we call a rational $`d`$-cone $`𝒦=𝐯+_{i=1}^d_0𝐰_i`$ *unimodular* if $`𝐰_1,\mathrm{},𝐰_d^d`$ generate the integer lattice $`^d`$. The significance of a unimodular cone $`𝒦`$ for us is that its fundamental (half-open) parallelepiped contains precisely one integer point $`𝐩`$, and so the generating function of $`𝒦`$ has a very simple and short form $$\sigma _𝒦(x)=\frac{x^𝐩}{\left(1x^{𝐰_1}\right)\mathrm{}\left(1x^{𝐰_d}\right)}.$$ In fact, the description length of this is proportional to the description of the cone $`𝒦`$. ###### Theorem 17 (Barvinok). For fixed dimension $`d`$, the generating function $`\sigma _𝒦`$ for any rational cone $`𝒦`$ in $`^d`$ can be decomposed into generating functions of unimodular cones in polynomial time; that is, there is a polynomial-time algorithm and (polynomially many) unimodular cones $`𝒦_j`$ such that $`\sigma _𝒦(x)=_jϵ_j\sigma _{𝒦_j}(x)`$, where $`ϵ_j\{\pm 1\}`$. Here *polynomial time* refers to the input data of $`𝒦`$, that is, the algorithm runs in time polynomial in the input length of, say, the halfspace description of $`𝒦`$. Brion’s Formula implies that an identical complexity statement can be made about the generating function $`\sigma _𝒫(x)`$ for any rational polytope $`𝒫`$. From here it is a short step (which nevertheless needs some justification) to see that one can *count* integer points in a rational polytope in polynomial time. We illustrate Barvinok’s short signed decomposition for the cone $`𝒦:=(0,0)+_0(1,0)+_0(1,4)`$, ignoring cones of smaller dimension. $`=`$ $``$ While $`𝒦`$ is the difference of two unimodular cones, it has a unique decomposition as a sum of four unimodular cones. $`=`$ $`+`$ $`+`$ $`+`$ In general the cone $`(0,0)+_0(1,0)+_0(1,n)`$ is the difference of two unimodular cones, but it has a unique decomposition into $`n`$ unimodular cones. Arguably the most famous consequence of Barvinok’s Theorem applies to *Ehrhart quasipolynomials*—the counting functions $`L_𝒫(t):=\mathrm{\#}\left(t𝒫^d\right)`$ in the positve-integer variable $`t`$ for a given rational polytope $`𝒫`$. One can show that the generating function $`_{t1}L_𝒫(t)x^t`$ is a rational function, and Barvinok’s Theorem implies that this rational function can be computed in polynomial time. Barvinok’s algorithm has been implemented in the software packages barvinok and LattE . The method of irrational decomposition has also been implemented in LattE, considerably improving its performance . ## Acknowledgments Research of Haase supported in part by NSF grant DMS-0200740 and DFG Emmy Noether fellowship. Research of Sottile supported in part by the Clay Mathematical nstitute and NSF CAREER grant DMS-0538734.
warning/0506/hep-ph0506148.html
ar5iv
text
# Study of the Fundamental Structure of Matter with an Electron-Ion Collider ## 1 Introduction Understanding the fundamental structure of matter is one of the central goals of scientific research. In the closing decades of the twentieth century, physicists developed a beautiful theory, Quantum Chromodynamics (QCD), which explains all of strongly interacting matter in terms of point-like quarks interacting by the exchange of gauge bosons, known as gluons. The gluons of QCD, unlike the photons of QED, can interact with each other. The color force which governs the interaction of quarks and gluons is responsible for more than 99% of the observable mass in the physical universe and explains the structure of nucleons and their composite structures, atomic nuclei, as well as astrophysical objects such as neutron stars. During the last 30 years, experiments have verified QCD quantitatively in collisions involving a very large momentum exchange between the participants. These collisions occur over very short distances much smaller than the size of the proton. In these experiments, the confined quarks and gluons act as if they are nearly free pointlike particles and exhibit many properties that are predicted by perturbative QCD (pQCD). This experimental phenomenon was first discovered in deeply inelastic scattering (DIS) experiments of electrons off nucleons. The discovery resulted in the 1990 Nobel Prize in Physics being awarded to Friedman, Kendall and Taylor. The phenomenon, that quarks and gluons are quasi-free at short distances, follows from a fundamental property of QCD known as asymptotic freedom. Gross, Politzer and Wilczek, who first identified and understood this unique characteristic of QCD were awarded the 2004 Nobel Prize in Physics. When the interaction distance between the quarks and gluons becomes comparable to or larger than the typical size of hadrons, the fundamental constituents of the nucleon are no longer free. They are confined by the strong force that does not allow for the observation of any “colored” object. In this strong coupling QCD regime, where most hadronic matter exists, the symmetries of the underlying quark-gluon theory are hidden, and QCD computations in terms of the dynamical properties of quarks and gluons are difficult. A major effort is underway worldwide to carry out ab initio QCD calculations in the strong QCD regime using Monte-Carlo simulations on large scale computers. The experimental underpinnings for QCD are derived from decades of work at the CERN, DESY, Fermilab and SLAC accelerator facilities. Some highlights include the determination of the nucleon quark momentum and spin distributions and the nucleon gluon momentum distribution, the verification of the QCD prediction for the running of the strong coupling constant $`\alpha _s`$, the discovery of jets, and the discovery that quark and gluon momentum distributions in a nucleus differ from those in a free nucleon. However, thirty years after QCD has been established as the Standard Model of the strong force, and despite impressive progress made in the intervening decades, understanding how QCD works in detail remains one of the outstanding issues in physics. Some crucial open questions that need to be addressed are listed below. * What is the gluon momentum distribution in the atomic nucleus? QCD tells us that the nucleon is primarily made up of specks of matter (quarks) bound by tremendously powerful gluon fields. Thus atomic nuclei are primarily composed of glue. Very little is known about the gluon momentum distribution in a nucleus. Determining these gluon distributions is therefore a fundamental measurement of high priority. This quantity is also essential for an understanding of other important questions in hadronic physics. For example, the interpretation of experiments searching for a deconfined quark-gluon state in relativistic heavy ion collisions is dependent on the knowledge of the initial quark and gluon configuration in a heavy nucleus. This will be especially true for heavy ion experiments at the Large Hadron Collider (LHC) at CERN. Further, there are predictions that gluonic matter at high parton densities has novel properties that can be probed in hard scattering experiments on nuclear targets. Hints of the existence of this state may have been seen in Deuteron-Gold experiments at the Relativistic Heavy Ion Collider (RHIC) at Brookhaven. * How is the spin structure of the nucleon understood to arise from the quark and gluon constituents? High energy spin-dependent lepton scattering experiments from polarized nucleon targets have produced surprising results. The spins of the quarks account for only about $`20\%`$ of the spin of the proton. The contribution of the gluons may be large. Dramatic effects are predicted for measurements beyond the capability of any existing accelerator. There are hints from other experiments that the contribution of orbital angular momentum may be large. * Testing QCD It is imperative to continue to subject QCD to stringent tests because there is so much about the theory that remains a mystery. QCD can be tested in two ways: one is by precision measurements, and the other is by looking for novel physics which is sensitive to the confining properties of the theory. Both of these can be achieved at a high luminosity lepton-ion collider with a detector that has a wide rapidity and angular coverage. An example of precision physics is the Bjorken Sum Rule in spin-dependent lepton scattering from a polarized nucleon. This fundamental sum rule relates inclusive spin-dependent lepton scattering to the ratio of axial to vector coupling constants in neutron $`\beta `$-decay. Present experiments test it to about $`\pm 10\%`$: it would be highly desirable to push these tests to about $`\pm `$1%. Further, with lattice QCD expected to make substantial progress in the ability to make ab initio QCD calculations during the next decade, precise measurements of the calculable observables will be required. An example of a physics measurement sensitive to confinement is hard diffraction, where large mass final states are formed with large “color-less” gaps in rapidity separating them from the hadron or nucleus. At the Hadron Electron Ring Accelerator (HERA) at DESY, roughly 10% of events are of this nature. The origins of these rapidity gaps, which must be intimately related to the confining properties of the theory, can be better understood with detectors that are able to provide detailed maps of the structure of events in DIS. This article motivates and describes the next generation accelerator required by nuclear and particle physicists to study and test QCD, namely a polarized lepton-ion collider. The basic characteristics of the collider are motivated as follows: * lepton beam The lepton probe employs the best understood interaction in nature (QED) to study hadron structure. Electrons and positrons couple directly to the quarks. The experimental conditions which maximize sensitivity to valence and sea quarks as well as probe gluons are well understood. Further, the availability of both positron and electron beams will enable experiments that are sensitive to the exchange of the parity violating Z and W-bosons. * range of center-of-mass (CM) energies To cleanly interact with quarks, a minimum center-of-mass (CM) energy of about 10 GeV is required. To explore and utilize the powerful $`Q^2`$ evolution equations of QCD, CM energies of order 100 GeV are desirable. This consideration strongly motivates the collider geometry. * high luminosity The QED interaction between the lepton probe and the hadron target is relatively weak. Thus precise and definitive measurements demand a high collision luminosity of order $`10^{33}`$ nucleons cm<sup>-2</sup> s<sup>-1</sup>. * polarized beams Polarized lepton and nucleon beams are essential to address the central question of the spin structure of the nucleon. Both polarized proton and neutron (effectively polarized <sup>2</sup>H or <sup>3</sup>He) are required for tests of the fundamental Bjorken Sum Rule. The polarization direction of at least one of the beams must be reversible on a rapid timescale to minimize systematic uncertainties. * nuclear beams Light nuclear targets are useful for probing the spin and flavor content of parton distributions. Heavy nuclei are essential for experiments probing the behavior of quarks and gluons in the nuclear medium. * detector considerations The collider geometry has a significant advantage over fixed-target experiments at high energy because it makes feasible the detection of complete final-states. A central collider detector with momentum and energy measurements and particle identification for both leptons and hadrons will be essential for many experiments. Special purpose detectors that provide wide angular and rapidity coverage will be essential for several specific measurements. These considerations constrain the design parameters of the collider to be a 5 to 10 GeV energy electron (or positron) beam colliding with a nucleon beam of energy 25 GeV to 250 GeV. The collider is anticipated to deliver nuclear beams of energies ranging from $`20100`$ GeV/nucleon. The lepton and nucleon beams must be highly polarized and the collision luminosity must be of order $`10^{33}`$ nucleons cm<sup>-2</sup> s<sup>-1</sup>. The proposed eRHIC design (described in section 4) realizes the required specifications in a cost effective and timely way by using the existing RHIC facility at BNL. The characteristics of eRHIC are well beyond the capability of any existing accelerator, as is clear from Fig. 1. By delivering high energies to the collision, the collider provides an increased range for investigating quarks and gluons with small momentum fraction ($`x`$) and for studying their behavior over a wide range of momentum transfers ($`Q^2`$). In deeply inelastic scattering, the accessible values of the Bjorken variable $`x`$ (defined in section 2) are limited by the available CM energy. For example, collisions between a 10 GeV lepton beam and nuclear beams of 100 GeV/nucleon provide access to values of $`x`$ as small as $`3\times 10^4`$ for $`Q^21`$ GeV<sup>2</sup>. In a fixed-target configuration, a 2.1 TeV lepton beam would be required to produce the same CM energy. Figure 2 shows the $`x`$-$`Q^2`$ range possible with the proposed eRHIC machine and compares that range to the currently explored kinematic region. In this article, the scientific case and accelerator design for a new facility to study the fundamental quark and gluon structure of strongly interacting matter is presented. Section 2 describes the current understanding of the quark and gluon structure of hadrons and nuclei. Section 3 presents highlights of the scientific opportunities available with a lepton-ion collider. Section 4 describes the accelerator design effort and section 5 describes the interaction region and eRHIC detector design. ## 2 Status of the Exploration of the Partonic Structure of Hadrons and Nuclei This section will summarize our current understanding of the partonic structure of hadrons and nuclei in QCD, accumulated during the past three decades from a variety of deeply inelastic and hadronic scattering experiments. We will also comment on what new information may become available from DIS as well as from RHIC and other experimental facilities around the world before a future electron-ion collider starts taking data. We will outline the status of our knowledge on i) the parton distributions in nucleons, ii) spin and flavor distributions in the nucleon, iii) nuclear modifications to the inclusive nucleon distributions such as the European Muon Collaboration (EMC) effect and quark and gluon shadowing, iv) color coherent phenomena in nuclei that probe the space-time structure of QCD such as color transparency and opacity, partonic energy loss and the $`p_T`$ broadening of partons in media. In each case, we will outline the most important remaining questions and challenges. These will be addressed further in Section 3. ### 2.1 Deeply-inelastic scattering The cross-section for the inclusive deeply inelastic scattering (DIS) process shown in Fig. 3 can be written as a product of the leptonic tensor $`_{\mu \nu }`$ and the hadronic tensor $`𝒲^{\mu \nu }`$ as $$\frac{d^2\sigma }{dxdy}_{\mu \nu }(k,q,s)𝒲^{\mu \nu }(P,q,S),$$ (1) where one defines the Lorentz invariant scalars, the famous Bjorken variable $`x=q^2/2Pq`$, and $`y=Pq/Pk`$. Note that as illustrated in Fig. 3, $`k`$ ($`k^{}`$) is the 4-momentum of the incoming (outgoing) electron, $`P`$ is the 4-momentum of the incoming hadron, and $`q=kk^{}`$ is the 4-momentum of the virtual photon. The center of mass energy squared is $`s=(P+k)^2`$. From these invariants, one can deduce simply that $`xyQ^2/s`$, where $`Q^2=q^2>0`$. The hadronic tensor can be written in full generality as $`𝒲^{\mu \nu }(P,q,S)={\displaystyle \frac{1}{4\pi }}{\displaystyle d^4z\mathrm{e}^{iqz}P,S|[𝒥_\mu (z),𝒥_\nu (0)]|P,S}=g^{\mu \nu }F_1(x,Q^2)`$ $`+{\displaystyle \frac{P^\mu P^\nu }{Pq}}F_2(x,Q^2)i\epsilon ^{\mu \nu \rho \sigma }{\displaystyle \frac{q_\rho P_\sigma }{2Pq}}F_3(x,Q^2)+i\epsilon ^{\mu \nu \rho \sigma }q_\rho [{\displaystyle \frac{S_\sigma }{Pq}}g_1(x,Q^2)`$ $`+{\displaystyle \frac{S_\sigma (Pq)P_\sigma (Sq)}{(Pq)^2}}g_2(x,Q^2)]+[{\displaystyle \frac{P^\mu S^\nu +S^\mu P^\nu }{2Pq}}{\displaystyle \frac{Sq}{(Pq)^2}}P^\mu P^\nu ]g_3(x,Q^2)`$ $`+{\displaystyle \frac{Sq}{(Pq)^2}}P^\mu P^\nu g_4(x,Q^2){\displaystyle \frac{Sq}{Pq}}g^{\mu \nu }g_5(x,Q^2).`$ (2) The $`F_i`$ are referred to as the “unpolarized” structure functions, whereas the $`g_i`$ are the “spin-dependent” ones, because their associated tensors depend on the nucleon spin vector $`S^\mu `$. Note that parity-violating interactions mediated by electroweak boson exchange are required for $`F_3,g_3,g_4,g_5`$ to contribute. Inserting Eq. 2 and the straightforwardly calculated leptonic tensor into Eq. 1, one obtains the DIS cross section in terms of the structure functions. If one averages over the hadronic spins and restricts oneself to parity conserving (for $`Q^2M_Z^2`$) electron-nucleon scattering alone, one finds the simple expression $`{\displaystyle \frac{d^2\sigma }{dxdQ^2}}={\displaystyle \frac{2\pi \alpha _{\mathrm{em}}^2}{Q^4}}\left[\left(1+(1y)^2\right)F_2(x,Q^2)y^2F_L(x,Q^2)\right].`$ (3) Here, $`\alpha _{\mathrm{em}}`$ is the coupling constant of Quantum Electrodynamics and $`F_L`$ is the “longitudinal” structure function, defined by the relation $`F_L=F_22xF_1`$. In the leading logarithmic approximation of QCD the measured structure function $`F_2(x,Q^2)`$ can be written as $`F_2(x,Q^2)={\displaystyle \underset{q=u,d,s,c,b,t}{}}e_q^2\left(xq(x,Q^2)+x\overline{q}(x,Q^2)\right),`$ (4) where $`q(x,Q^2)`$ ($`\overline{q}(x,Q^2))`$ is the probability density for finding a quark (anti-quark) with momentum fraction $`x`$ at a momentum resolution scale $`Q^2`$; $`e_q`$ is the quark charge. In the simple parton model one has Bjorken scaling, $`F_2(x,Q^2)F_2(x)`$. The “scaling violations” seen in the $`Q^2`$-dependence of $`F_2(x,Q^2)`$ arise from the fact that QCD is not a scale invariant theory and has an intrinsic scale $`\mathrm{\Lambda }_{\mathrm{QCD}}200`$ MeV. They are only logarithmic in the Bjorken limit of $`Q^2\mathrm{}`$ and $`s\mathrm{}`$ with $`xQ^2/s`$ fixed. As one moves away from the asymptotic regime, the scaling violations become significant. They can be quantitatively computed in QCD perturbation theory using for example the machinery of the operator product expansion and the renormalization group. The result is most conveniently summarized by the Dokshitzer-Gribov-Lipatov-Altarelli-Parisi (DGLAP) evolution equations for the parton densities (?, ?): $$\frac{d}{d\mathrm{ln}Q^2}\left(\begin{array}{c}q\\ g\end{array}\right)(x,Q^2)=\left(\begin{array}{cc}P_{qq}(\alpha _s,x)& P_{qg}(\alpha _s,x)\\ P_{gq}(\alpha _s,x)& P_{gg}(\alpha _s,x)\end{array}\right)\left(\begin{array}{c}q\\ g\end{array}\right)(x,Q^2),$$ (5) where $``$ denotes a convolution, and the $`P_{ij}`$ are known as “splitting functions” (?) and are evaluated in QCD perturbation theory. They are now known to three-loop accuracy (?). The evolution of the quark densities $`q`$, $`\overline{q}`$ involves the gluon density $`g(x,Q^2)`$. The physical picture behind evolution is the fact that the virtuality $`Q^2`$ of the probe sets a resolution scale for the partons, so that a change in $`Q^2`$ corresponds to a change in the parton state seen. The strategy is then to parameterize the parton distributions at some initial scale $`Q^2=Q_0^2`$, and to determine the parameters by evolving the parton densities to (usually, larger) $`Q^2`$ and by comparing to experimental data for $`F_2(x,Q^2)`$. The pioneering DIS experiments, which first measured Bjorken scaling of $`F_2`$, were performed at SLAC (?). However, because of the (relatively) small energies, these experiments were limited to the region of $`x0.1`$. With the intense muon beams of CERN and Fermilab, with energies in excess of 100 GeV, the DIS cross-section of the proton was measured down to and below $`x10^3`$ (?). In the 1990’s, the HERA collider at DESY extended the DIS cross-section of the proton to below $`x=10^4`$ (?, ?). The current experimental determination of $`F_2^{\mathrm{proton}}(x,Q^2)`$ extends over 4 orders of magnitude in $`x`$ and $`Q^2`$. This is shown in Fig. 4. The left panel in Fig. 4 shows next-to-leading order (NLO) QCD global fits by the ZEUS and H1 detector collaborations at HERA to $`F_2`$ as a function of $`Q^2`$ for the world DIS data. The data and the QCD fit are in excellent agreement over a wide range in $`x`$ and $`Q^2`$. In the right panel of Fig. 4, the $`x`$ dependence of $`F_2`$ is shown for different bins in $`Q^2`$. The rapid rise in $`F_2`$ with decreasing $`x`$ reflects the sizeable contribution from the sea quark distribution at small $`x`$. In Fig. 5 we show the valence up and down quark distributions as well as the gluon and sea quark distributions extracted by the H1 and ZEUS collaborations as functions of $`x`$ for fixed $`Q^2=10`$ GeV<sup>2</sup>. The valence parton distributions are mainly distributed at large $`x`$ whereas the glue and sea quark distributions dominate hugely at small $`x`$. Indeed, the gluon and sea quark distributions are divided by a factor of 20 to ensure they can be shown on the scale of the plot. Already at $`x0.1`$, the gluon distribution is nearly a factor of two greater than the sum of the up and down quark valence distributions. As follows from Eq. 5, the gluon distribution in DIS may be extracted from scaling violations of $`F_2`$: $`xg(x,Q^2)\frac{F_2(x,Q^2)}{\mathrm{ln}Q^2}`$. As one goes to low $`Q^2`$, $`xg(x,Q^2)`$ becomes small, and some analyses find a preference for a negative gluon distribution at low $`x`$, modulo statistical and systematic uncertainties (?, ?). This is in principle not a problem in QCD beyond leading order. However, the resulting longitudinal structure function $`F_L`$ also comes out close to zero or even negative for $`Q^22`$ GeV<sup>2</sup> <sup>1</sup><sup>1</sup>1The leading twist expression for $`F_L`$ is simply related to $`\alpha _Sxg(x,Q^2)`$., which is unphysical because $`F_L`$ is a positive-definite quantity. A likely explanation for this finding is that contributions to $`F_L`$ that are suppressed by inverse powers of $`Q^2`$ are playing a significant role at these values of $`Q^2`$ (?). These contributions are commonly referred to as higher twist effects. It has been shown recently (?) that the HERA data on the virtual photon-proton cross-section ($`\sigma ^{\gamma ^{}p}=4\pi ^2\alpha _{\mathrm{em}}F_2(x,Q^2)/Q^2`$), for all $`x10^2`$ and $`0.045Q^2<450`$ GeV<sup>2</sup>, exhibit the phenomenon of “geometrical scaling” shown in Fig. 6. The data are shown to scale as a function of $`\tau =Q^2/Q_s^2`$, where $`Q_s^2(x)=Q_0^2(x_0/x)^\lambda `$ with $`Q_0^2=1`$ GeV<sup>2</sup>, $`x_0=310^4`$ and $`\lambda 0.3`$. The scale $`Q_s^2`$ is called the saturation scale. Geometrical scaling, although very general, is realized in a simple model, the Golec-Biernat-Wüsthoff model which includes all twist contributions (?). The model (and variants) provide a phenomenological description of the HERA data on diffractive cross-sections and inclusive vector meson production (?, ?, ?, ?, ?). The saturation scale and geometrical scaling will be discussed further in Section 3. ### 2.2 Spin structure of the nucleon #### 2.2.1 What we have learned from polarized DIS Spin physics has played a prominent role in QCD for several decades. The field has been driven by the successful experimental program of polarized deeply-inelastic lepton-nucleon scattering at SLAC, CERN, DESY and the Jefferson Laboratory (?). A main focus has been on measurements with longitudinally polarized lepton beam and target. For leptons with helicity $`\lambda `$ scattering off nucleons polarized parallel or antiparallel to the lepton direction, one has (?) $$\frac{d^2\sigma ^{\lambda ,}}{dxdQ^2}\frac{d^2\sigma ^{\lambda ,}}{dxdQ^2}C(G_v,G_a,\lambda )\left[\lambda xy(2y)g_1+(1y)g_4+xy^2g_5\right],$$ (6) where $`C(G_v,G_a,\lambda )`$ are factors depending on the vector and axial couplings of the lepton to the exchanged gauge boson. The terms involving $`g_4`$ and $`g_5`$ in Eq. 6 are associated with $`Z`$ and $`W`$ exchange in the DIS process and violate parity. In the fixed-target regime, pure-photon exchange strongly dominates, and scattering off a longitudinally polarized target determines $`g_1`$. Figure 7 (left) shows a recent compilation (?) of the world data on $`g_1(x,Q^2)`$, for proton, deuteron, and neutron targets. Roughly speaking, $`g_1`$ is known about as well now as the unpolarized $`F_2`$ was in the mid-eighties, prior to HERA. Figure 7 (right) shows the measured $`Q^2`$-dependence of $`g_1`$; the predicted scaling violations are visible in the data. In leading order of QCD, $`g_1`$ can be written as $$g_1(x,Q^2)=\frac{1}{2}\underset{q}{}e_q^2\left[\mathrm{\Delta }q(x,Q^2)+\mathrm{\Delta }\overline{q}(x,Q^2)\right],$$ (7) where $$\mathrm{\Delta }qq_{}^{}q_{}^{}(q=u,d,s,\mathrm{}),$$ (8) $`q_{}^{}`$ ($`q_{}^{}`$) denoting the number density of quarks of same (opposite) helicity as the nucleon. Clearly, the $`\mathrm{\Delta }q(x,Q^2),\mathrm{\Delta }\overline{q}(x,Q^2)`$ contain information on the nucleon spin structure. Also in the spin-dependent case, QCD predicts $`Q^2`$-dependence of the densities. The associated evolution equations have the same form as Eq. 5, but with polarized splitting functions (?, ?, ?). Also, the spin-dependent gluon density $`\mathrm{\Delta }g`$, defined in analogy with Equation 8, appears. The results of a recent QCD analysis (?) of the data for $`g_1(x,Q^2)`$ in terms of the polarized parton densities are shown in Fig. 8. The shaded bands in the figure give estimates of how well we know the distributions so far. As can be seen, the valence densities are fairly well known and the sea quark densities to some lesser extent. This analysis (?) assumes flavor-SU(3) symmetry for the sea quarks; the actual uncertainties in the individual sea distributions are much larger. Finally, Fig. 8 also shows that we know very little about the polarized gluon density. A tendency toward a positive $`\mathrm{\Delta }g`$ is seen. It is not surprising that the uncertainty in $`\mathrm{\Delta }g`$ is still large: at LO, $`\mathrm{\Delta }g`$ enters only through the $`Q^2`$-evolution of the structure function $`g_1`$. Because all polarized DIS experiments thus far have been with fixed targets, the lever arm in $`Q^2`$ has been limited. This is also seen in a comparison of Fig. 7 with Fig. 4. A particular focus in the analysis of $`g_1`$ has been on the integral $`\mathrm{\Gamma }_1(Q^2)_0^1g_1(x,Q^2)𝑑x`$. Ignoring QCD corrections, one has from Eq. 7: $$\mathrm{\Gamma }_1=\frac{1}{12}\mathrm{\Delta }𝒜_3+\frac{1}{36}\mathrm{\Delta }𝒜_8+\frac{1}{9}\mathrm{\Delta }\mathrm{\Sigma },$$ (9) where $`\mathrm{\Delta }\mathrm{\Sigma }`$ $`=`$ $`\mathrm{\Delta }𝒰+\mathrm{\Delta }\overline{𝒰}+\mathrm{\Delta }𝒟+\mathrm{\Delta }\overline{𝒟}+\mathrm{\Delta }𝒮+\mathrm{\Delta }\overline{𝒮},`$ $`\mathrm{\Delta }𝒜_3`$ $`=`$ $`\mathrm{\Delta }𝒰+\mathrm{\Delta }\overline{𝒰}\mathrm{\Delta }𝒟\mathrm{\Delta }\overline{𝒟},`$ $`\mathrm{\Delta }𝒜_8`$ $`=`$ $`\mathrm{\Delta }𝒰+\mathrm{\Delta }\overline{𝒰}+\mathrm{\Delta }𝒟+\mathrm{\Delta }\overline{𝒟}2\left(\mathrm{\Delta }𝒮+\mathrm{\Delta }\overline{𝒮}\right),`$ (10) with $`\mathrm{\Delta }𝒬=_0^1\mathrm{\Delta }q(x,Q^2)𝑑x`$, which does not evolve with $`Q^2`$ at lowest order. The flavor non-singlet combinations $`\mathrm{\Delta }𝒜_i`$ turn out to be proportional to the nucleon matrix elements of the quark non-singlet axial currents, $`P,S|\overline{q}\gamma ^\mu \gamma ^5\lambda _iq|P,S`$. Such currents typically occur in weak interactions, and by SU(3) rotations one may relate the matrix elements to the $`\beta `$-decay parameters $`F,D`$ of the baryon octet (?, ?). One finds $`\mathrm{\Delta }𝒜_3=F+D=g_A=1.267`$ and $`\mathrm{\Delta }𝒜_8=3FD0.58`$. The first of these remarkable connections between hadronic and DIS physics corresponds to the famous Bjorken sum rule (?), $$\mathrm{\Gamma }_1^p\mathrm{\Gamma }_1^n=\frac{1}{6}\mathrm{\Delta }𝒜_3\left[1+𝒪(\alpha _s)\right]=\frac{1}{6}g_A\left[1+𝒪(\alpha _s)\right],$$ (11) where the superscripts $`p`$ and $`n`$ denote the proton and neutron respectively. The sum rule has been verified experimentally with about $`10\%`$ accuracy (?). The QCD corrections indicated in Equation 11 are known (?) through $`𝒪(\alpha _s^3)`$. Assuming the validity of the sum rule, it can be used for a rather precise determination of the strong coupling constant (?). Determining $`\mathrm{\Gamma }_1`$ from the polarized-DIS data, and using the information from $`\beta `$-decays on $`\mathrm{\Delta }𝒜_3`$ and $`\mathrm{\Delta }𝒜_8`$ as additional input, one may determine $`\mathrm{\Delta }\mathrm{\Sigma }`$. This quantity is of particular importance because it measures twice the quark spin contribution to the proton spin. The analysis reveals a small value $`\mathrm{\Delta }\mathrm{\Sigma }0.2`$. The experimental finding that the quarks carry only about 20$`\%`$ of the proton spin has been one of the most remarkable results in the exploration of the structure of the nucleon. Even though the identification of nucleon with parton helicity is not a prediction of QCD (perturbative or otherwise) the result came as a major surprise. It has sparked tremendous theoretical activity and has also been the motivation behind a number of dedicated experiments in QCD spin physics, aimed at further unraveling the nucleon spin. A small value for $`\mathrm{\Delta }\mathrm{\Sigma }`$ also implies a sizable negative strange quark polarization in the nucleon, $`\mathrm{\Delta }𝒮+\mathrm{\Delta }\overline{𝒮}0.12`$. It would be desirable to have independent experimental information on this quantity, to eliminate the uncertainty in the value for $`\mathrm{\Delta }\mathrm{\Sigma }`$ due to SU(3) breaking effects in the determination of $`\mathrm{\Delta }𝒜_8`$ from baryon $`\beta `$ decays (?). More generally, considering Fig. 8, more information is needed on the polarized sea quark distribution functions and their flavor decomposition. Such knowledge is also very interesting for comparisons to model calculations of nucleon structure. For example, there have been a number of predictions (?) for the $`\mathrm{\Delta }\overline{u}\mathrm{\Delta }\overline{d}`$. Progress toward achieving a full flavor separation of the nucleon sea has been made recently, through semi-inclusive measurements in DIS (SIDIS) (?, ?). Inclusive DIS via photon exchange only gives access to the combinations $`\mathrm{\Delta }q+\mathrm{\Delta }\overline{q}`$, as is evident from Equation 7. If one detects, however, a hadron in the final state, the spin-dependent structure function becomes $$g_1^h(x,z)=\frac{1}{2}\underset{q}{}e_q^2\left[\mathrm{\Delta }q(x)D_q^h(z)+\mathrm{\Delta }\overline{q}(x)D_{\overline{q}}^h(z)\right].$$ (12) Here, the $`D_i^h(z)`$ are fragmentation functions, with $`z=E^h/\nu `$, where $`E^h`$ is the energy of the produced hadron and $`\nu `$ the energy of the virtual photon in the Lab frame. Fig. 9 shows the latest results on the flavor separation by the HERMES collaboration at HERA (?). Uncertainties are still fairly large; unfortunately, no further improvements in statistics are expected from HERMES. The results are not inconsistent with the large negative polarization of $`\mathrm{\Delta }\overline{u}=\mathrm{\Delta }\overline{d}=\mathrm{\Delta }\overline{s}`$ in the sea that has been implemented in many determinations of polarized parton distributions from inclusive DIS data (see, e.g., the curves in Fig. 8). On the other hand, there is no evidence for a large negative strange quark polarization. The results have sparked much renewed theory activity on SIDIS (?). We note that at RHIC $`W^\pm `$ production will be used to determine $`\mathrm{\Delta }u,\mathrm{\Delta }\overline{u},\mathrm{\Delta }d,\mathrm{\Delta }\overline{d}`$ with good precision, exploiting the parity-violating couplings of the $`W`$ to left-handed quarks and right-handed antiquarks (?). Comparisons of such data taken at much higher scales with those from SIDIS will be extremely interesting. A measurement of $`\mathrm{\Gamma }_1`$ obviously relies on an estimate of the contribution to the integral from $`x`$ outside the measured region. The extrapolation to small $`x`$ constitutes one main uncertainty in the value of $`\mathrm{\Delta }\mathrm{\Sigma }`$. As can be seen from Fig. 7, there is not much information on $`g_1(x,Q^2)`$ at $`x<0.003`$. In addition, the data points at the smaller $`x`$ also have $`Q^2`$ values that are below the DIS regime, making it conceivable that the “higher-twist” contributions to $`g_1(x,Q^2)`$ are important and contaminate the extraction of $`\mathrm{\Delta }\mathrm{\Sigma }`$. About half of the data points shown in Fig. 7 are from the region $`Q^24`$ GeV<sup>2</sup>, $`W^2=Q^2(1x)/x10`$ GeV<sup>2</sup>, which in the unpolarized case is usually excluded in analyses of parton distribution functions. Clearly, measurements of polarized DIS and SIDIS at smaller $`x`$, as well as at presently available $`x`$, but higher $`Q^2`$, will be vital for arriving at a definitive understanding of the polarized quark distributions, and of $`\mathrm{\Delta }\mathrm{\Sigma }`$ in particular. #### 2.2.2 Contributors to the nucleon spin The partons in the nucleon have to provide the nucleon spin. When formulating a “proton spin sum rule” one has in mind the expectation value of the angular momentum operator (?, ?), $$\frac{1}{2}=P,1/2|\widehat{J}_3|P,1/2=P,1/2|d^3\left[\stackrel{}{x}\times \stackrel{}{T}\right]_3|P,1/2,$$ (13) where $`T^i𝒯^{0i}`$ with $`𝒯`$ the QCD energy-momentum tensor. Expressing the operator in terms of quark and gluon operators, one may write: $$\frac{1}{2}=\frac{1}{2}\mathrm{\Delta }\mathrm{\Sigma }+\mathrm{\Delta }G(Q^2)+L_q(Q^2)+L_g(Q^2),$$ (14) where $`\mathrm{\Delta }G(Q^2)=_0^1\mathrm{\Delta }g(x,Q^2)`$ is the gluon spin contribution and the $`L_{q,g}`$ correspond to orbital angular momenta of quarks and gluons. Unlike $`\mathrm{\Delta }\mathrm{\Sigma }`$, $`\mathrm{\Delta }G`$ and $`L_{q,g}`$ depend on the resolution scale $`Q^2`$ already at lowest order in evolution. The small size of the quark spin contribution implies that we must look elsewhere for the proton’s spin: sizable contributions to the nucleon spin should come from $`\mathrm{\Delta }G`$ and/or $`L_{q,g}`$. Several current experiments are dedicated to a direct determination of $`\mathrm{\Delta }g(x,Q^2)`$. High-transverse momentum jet, hadron, and photon final states in polarized $`pp`$ scattering at RHIC offer the best possibilities (?). For example, direct access to $`\mathrm{\Delta }g`$ is provided by the spin asymmetry for the reaction $`pp\gamma X`$, owing to the presence of the QCD Compton process $`qg\gamma q`$. The Spin Muon Collaboration (SMC) and COMPASS fixed-target experiments at CERN, and the HERMES experiment at DESY, access $`\mathrm{\Delta }g(x,Q^2)`$ in charm or high-$`p_T`$ hadron pair final states in photon-gluon fusion $`\gamma ^{}gq\overline{q}`$ (?). Additional precision measurements with well established techniques will be needed to determine the integral of the polarized gluon distribution, particularly at lower $`x`$. Orbital effects are the other candidate for contributions to the proton spin. Close analysis of the $`\stackrel{}{x}\times \stackrel{}{T}`$ matrix elements in Eq. 13 revealed (?) that they can be measured from a wider class of parton distribution functions, the so-called generalized parton distributions (GPD) (?). These take the general form $`p+\mathrm{\Delta }|𝒪_{q,g}|p`$, where $`𝒪_{q,g}`$ are suitable quark and gluon operators and $`\mathrm{\Delta }`$ is some momentum transfer. The latter is the reason that the GPDs are also referred to as “off-forward” distributions. The explicit factor $`\stackrel{}{x}`$ in Equation 13 forces one off the forward direction, simply because it requires a derivative with respect to momentum transfer. This is in analogy with the nucleon’s Pauli form factor. In fact, matrix elements of the above form interpolate between DIS structure functions and elastic form factors. To be more specific (?), the total (spin plus orbital) angular momentum contribution of a quark to the nucleon spin is given as (?) $$J_q=\frac{1}{2}\underset{\mathrm{\Delta }^20}{lim}𝑑xx\left[H_q(x,\xi ,\mathrm{\Delta }^2)+E_q(x,\xi ,\mathrm{\Delta }^2)\right].$$ (15) Here, $`\xi =\mathrm{\Delta }^+/P^+`$, where the light-cone momentum $`\mathrm{\Delta }^+\mathrm{\Delta }^0+\mathrm{\Delta }^z`$, and likewise for $`P^+`$. $`H_q,E_q`$ are defined as form factors of the matrix element $`𝑑y\mathrm{e}^{iyx}P^{}|\overline{\psi }_+(y)\psi _+\left(0\right)|P`$. $`H_q`$ reduces to the ordinary (forward) quark distribution in the limit $`\mathrm{\Delta }0`$, $`H_q(x,0,0)=q(x)`$, whereas the first moments (in $`x`$) of $`H_q`$ and $`E_q`$ give the quark’s contributions to the nucleon Dirac and Pauli form factors, respectively. In addition, Fourier transforms of $`H_q,E_q`$ with respect to the transverse components of the momentum transfer $`\mathrm{\Delta }`$ give information on the position space distributions of partons in the nucleon (?), for example: $$H_q(x,\xi =0,\stackrel{}{\mathrm{\Delta }}_{}^2)=d^2\stackrel{}{b}\mathrm{e}^{i\stackrel{}{\mathrm{\Delta }}_{}\stackrel{}{b}}q(x,b).$$ (16) $`q(x,b)`$ is the probability density for finding a quark with momentum fraction $`x`$ at transverse distance $`\stackrel{}{b}`$ from the center. It thus gives a transverse profile of the nucleon. GPDs, therefore, may give us remarkably deep new insight into the nucleon. The classic reaction for a measurement of the $`H_q,E_q`$ is “deeply virtual Compton scattering (DVCS)”, $`\gamma ^{}p\gamma p`$ (?). It is the theoretically best explored and understood reaction (?). Next-to-leading order calculations are available (?). The GPDs contribute to the reaction at amplitude level. The amplitude for DVCS interferes with that for the Bethe-Heitler process. The pure Bethe-Heitler part of the differential $`ep`$ cross section is calculable and can in principle be subtracted, provided it does not dominate too strongly. Such a subtraction has been performed in DVCS measurements at small $`x`$ by H1 and ZEUS (?). A different possibility to eliminate the Bethe-Heitler contribution is to take the difference of cross sections for opposite beam or target polarization. In both cases, contributions from Compton scattering and the Compton-Bethe-Heitler interference survive. The cleanest separation of these pieces can be achieved in experiments with lepton beams of either charge. Because the Compton contribution to the $`ep`$ amplitude is linear and the Bethe-Heitler contribution quadratic in the lepton charge, the interference term is projected out in the difference $`d\sigma (e^+p)d\sigma (e^{}p)`$ of cross sections, whereas it is absent in their sum. Both the “beam-spin” asymmetry $$\frac{d\sigma _+(e^{}p)d\sigma _{}(e^{}p)}{d\sigma _+(e^{}p)+d\sigma _{}(e^{}p)},$$ (17) where $`\pm `$ denote positive (negative) beam helicities, and the “beam-charge” asymmetry $$\frac{d\sigma (e^+p)d\sigma (e^{}p)}{d\sigma (e^+p)+d\sigma (e^{}p)}$$ (18) have been observed (?, ?, ?). Fig. 10 shows some of the results. Hard exclusive meson production, $`\gamma ^{}pMp`$, is another process that gives access to GPDs, and much activity has gone into this direction as well (?, ?). Both DVCS and exclusive meson production have their practical advantages and disadvantages. Real photon production is cleaner, but the price to be paid is an additional power of $`\alpha _{\mathrm{em}}`$. Meson production may be easier to detect; however, its amplitude is suppressed relatively by a power $`1/Q`$. The importance of using nucleon polarization in off-forward reactions is well established. There have also been first studies for DVCS off nuclei (?). Practical problems are the fact that GPDs depend on three variables (plus a scale in which they evolve), and that they appear in complicated convolutions with the partonic hard-scattering kernels. We are still far from the quantitative experimental surveys of DVCS and related processes that would allow us to work backwards to new insights into off-diagonal matrix elements and angular momentum. Nevertheless, a direction for the field has been set. #### 2.2.3 Transverse polarization In addition to the unpolarized and the helicity-dependent distributions, there is a third set of twist-2 parton distributions, namely transversity (?). In analogy with Equation 8 these distributions measure the net number (parallel minus antiparallel) of partons with transverse polarization in a transversely polarized nucleon: $$\delta q(x)=q_{}^{}(x)q_{}^{}(x).$$ (19) In a helicity basis (?), transversity corresponds to an interference of an amplitude in which a helicity-$`+`$ quark emerges from a helicity-$`+`$ nucleon, but is returned as a quark of negative helicity into a nucleon of negative helicity. This helicity-flip structure makes transversity a probe of chiral symmetry breaking in QCD (?). Perturbative-QCD interactions preserve chirality, and so the helicity flip must primarily come from soft non-perturbative interactions for which chiral symmetry is broken. The required helicity flip also precludes a gluon transversity distribution at leading twist (?). Measurements of transversity are not straightforward. Again the fact that perturbative interactions in the Standard Model do not change chirality (or, for massless quarks, helicity) means that inclusive DIS is not useful. Collins, however, showed (?) that properties of fragmentation might be exploited to obtain a “transversity polarimeter”: a pion produced in fragmentation will have some transverse momentum with respect to the fragmenting parent quark. There may then be a correlation of the form $`i\stackrel{}{S}_T(\stackrel{}{P}_\pi \times \stackrel{}{k}_{})`$ among the transverse spin $`\stackrel{}{S}_T`$ of the fragmenting quark, the pion momentum $`\stackrel{}{P}_\pi `$, and the transverse momentum $`\stackrel{}{k}_{}`$ of the quark relative to the pion. The fragmentation function associated with this correlation is denoted as $`H_1^{,q}(z)`$, the Collins function. If non-vanishing, the Collins function makes a leading-power (?, ?, ?) contribution to the single-spin asymmetry $`A_{}`$ in the reaction $`ep^{}e\pi X`$: $$A_{}|\stackrel{}{S}_T|\mathrm{sin}(\varphi +\varphi _S)\underset{q}{}e_q^2\delta q(x)H_1^{,q}(z),$$ (20) where $`\varphi `$ ($`\varphi _S`$) is the angle between the lepton plane and the $`(\gamma ^{}\pi )`$ plane (and the transverse target spin). As shown in Equation 20, this asymmetry would then allow access to transversity. If “intrinsic” transverse momentum in the fragmentation process can play a crucial role in the asymmetry for $`ep^{}e\pi X`$, a natural question is whether $`k_{}`$ in the initial state can be relevant as well. Sivers suggested (?) that the $`k_{}`$ distribution of a quark in a transversely polarized hadron could have an azimuthal asymmetry, $`\stackrel{}{S}_T(\stackrel{}{P}\times \stackrel{}{k}_{})`$. It was realized (?, ?) that the Wilson lines in the operators defining the Sivers function, required by gauge invariance, are crucial for the function to be non-vanishing. This intriguing discovery has been one of the most important theoretical developments in QCD spin physics in the past years. Another important aspect of the Sivers function is that it arises as an interference of wave functions with angular momenta $`J_z=\pm 1/2`$ and hence contains information on parton orbital angular momentum (?, ?), complementary to that obtainable from DVCS. Model calculations and phenomenological studies of the Sivers functions $`f_{1T}^{,q}`$ have been presented (?). It makes a contribution to $`ep^{}e\pi X`$ (?), $$A_{}|\stackrel{}{S}_T|\mathrm{sin}(\varphi \varphi _S)\underset{q}{}e_q^2f_{1T}^{,q}(x)D_q^\pi (z).$$ (21) This is in competition with the Collins function contribution, Equation 20; however, the azimuthal angular dependence is discernibly different. HERMES has completed a run with transverse polarization and performed an extraction of the contributions from the Sivers and Collins effects (?). There are also first results from COMPASS (?). First independent information on the Collins functions is now coming from Belle measurements in $`e^+e^{}`$ annihilation (?). The Collins and Sivers functions are also likely involved (?) in explanations of experimental observations of very large single-transverse spin asymmetries in $`pp`$ scattering (?), where none were expected. It was pointed out (?, ?) that comparisons of DIS results and results from $`p^{}p`$ scattering at RHIC will be particularly interesting: from the properties of the Wilson lines it follows that the Sivers functions violate universality of the distribution functions. For example, the Sivers functions relevant in DIS and in the Drell-Yan process should have opposite sign. This is a striking prediction awaiting experimental testing. ### 2.3 Nuclear modifications The nucleus is traditionally described as a collection of weakly bound nucleons confined in a potential created by their mutual interaction. It came as a surprise when the EMC experiment (?) uncovered a systematic nuclear dependence to the nuclear structure function $`F_2^A(x,Q^2)`$ in iron relative to that for Deuterium because the effect was as much as 20% for $`x0.5`$. This is significantly larger than the effect ($`<5`$%) due to the natural scale for nuclear effects given by the ratio of the binding energy per nucleon to the nucleon mass. Several dedicated fixed target experiments (?, ?, ?) confirmed the existence of the nuclear dependence observed by the EMC albeit with significant modifications of the original EMC results at small $`x`$. The upper part of Figure 11 shows an idealized version of the nuclear modification of the relative structure functions per nucleon. It is $`2/A`$ times the ratio of a measured nuclear structure function of nucleus A to that for Deuterium. The rise at the largest values of $`x`$ is ascribed to the nucleons’ Fermi momentum. The region above $`x0.2`$ is referred to as the EMC effect region. When $`x0.05`$, the nuclear ratio drops below one and the region is referred to as the nuclear shadowing region, whereas the region with the slight enhancement in between the shadowing and EMC effect regions is called the anti-shadowing region. The lower part of Figure 11 presents a sample of high precision data of ratios of structure functions over a broad range in $`A`$, $`x`$ and $`Q^2`$. We shall now discuss what is known about these regions, focusing in particular on the EMC effect and nuclear shadowing regions. #### 2.3.1 The EMC effect A review of the DIS data and various interpretations of the EMC effect since its discovery in the early 1980’s can be found in Ref. (?). A common interpretation of the EMC effect is based on models where inter-nucleon interactions at a wide range of inter-nucleon distances are mediated by meson exchanges. The traditional theory (?, ?) of nuclear interactions predicts a net increase in the distribution of virtual pions with increasing nuclear density relative to that of free nucleons. This is because meson interactions are attractive in nuclei. In these models, nuclear pions may carry about 5% of the total momentum to fit the EMC effect at $`x0.3`$. Each pion carries a light-cone fraction of about 0.2-0.3 of that for a nucleon. Sea anti-quarks belonging to these nuclear pions may scatter off a hard probe. Hence the predicted enhancement of the nuclear sea of 10% to 15% for $`x0.1`$–0.2 and for $`A40`$. The conventional view of nuclear binding is challenged by the constancy with $`A`$ of the anti-quark distribution extracted from the production of Drell-Yan pairs in proton-nucleus collisions at Fermilab (?). These data are shown in Fig. 12. No enhancement was observed at the level of 1% accuracy in the Drell-Yan experiments. The Drell-Yan data was also compared with and showed good agreement with the DIS EMC data for the $`F_2`$ ratio of Tin to Deuterium. Furthermore, first results (?) from the Jefferson Laboratory (TJNAF) experiment E91-003 indicate that there is no significant pion excess in the $`A(e,e^{},B)`$ reaction. (It has however been pointed out (?) that parameters of pion interactions in nuclei can be readily adjusted to reduce the pion excess to conform with the Drell-Yan data.) In addition, the energy excitation for the residual nuclear system also reduces the contribution of pions to the nuclear parton densities (?). Thus all of these observations suggest that pions may not contribute significantly to $`F_2^A`$ in the EMC region. The chiral quark-soliton model (?) is a phenomenologically successful model that for instance explains the difference in the anti-quark up and down distributions as a function of $`x`$ (?). Interestingly, it has been shown recently (?) to simultaneously provide a good description of both the EMC effect and the ratio of anti-quark distributions from Drell-Yan pairs. Recently, it has been argued that a key feature of the EMC effect, the factorization of the $`x`$ and $`A`$ dependence of the EMC ratio, can be understood in a model independent way in an effective field theory approach (?). Several joint leading order QCD analyses of the nuclear DIS and Drell-Yan data combined with the application of the baryon charge and momentum sum rules (?, ?, ?, ?) provide further information on the nuclear effects on parton densities in this kinematic region. These analyses indicate that the valence quark distribution in nuclei is enhanced at $`x0.1`$–0.2. Gluons in nuclei carry practically the same fraction of the momentum (within 1%) as in a free nucleon. If one assumes that gluon shadowing is similar to that for quarks, these analyses predict a significant enhancement of the gluon distribution in nuclei at $`x0.1`$–0.2 (?). A recent next-to-leading order (NLO) analysis of nuclear parton distributions (?) however finds that this gluon “anti-shadowing” is much smaller than in the LO analysis. #### 2.3.2 Nuclear shadowing Nuclear shadowing is the phenomenon, shown in Fig. 11, where the ratio of the nuclear electromagnetic structure function $`F_2^A`$ relative to $`A/2`$ times the Deuteron electromagnetic structure functions $`F_2^D`$ is less than unity for $`x0.05`$. Shadowing is greater for decreasing $`x`$ and with increasing nuclear size. For moderately small $`x`$, shadowing is observed to decrease slowly with increasing $`Q^2`$. Unfortunately, because $`x`$ and $`Q^2`$ are inversely correlated for fixed energies, much of the very small $`x`$ data ($`x10^3`$) is at very low values of $`Q^21`$ GeV<sup>2</sup>. In addition, as results (?) from the fixed target E665 experiment at Fermilab and the New Muon Collaboration (NMC) experiment at CERN shown in Fig. 11(b) suggest, good quality data exists only for $`x>4\times 10^3`$. At high $`Q^2`$, the shadowing of $`F_2^A`$ can be interpreted in terms of shadowing of quark and anti-quark distributions in nuclei at small $`x`$. Information on quark shadowing can also be obtained from proton-nucleus Drell-Yan experiments (?) and from neutrino-nucleus experiments–most recently from NuTeV at Fermilab (?). The phenomenon of shadowing has different interpretations depending on the frame in which we consider the space-time evolution of the scattering. Consider for instance the rest frame of a nucleus in $`\gamma `$-p/A scattering. The $`\gamma `$p cross-section is only 0.1mb for energies in excess of 2 GeV, corresponding to a mean-free-path of well over 100 fm in nuclear matter. However, although the high-energy $`\gamma `$A cross-section might be expected to be proportional to $`A`$, the observed increase in the cross-section is smaller than $`A`$ times the $`\gamma `$p cross-section. This is because the photon can fluctuate into a $`q\overline{q}`$– pair that has a cross-section typical of the strong interactions ($`20`$mb) and is absorbed readily (with a mean free path of $`3.5`$ fm). If the fluctuation persists over a length greater than the inter-nucleon separation distance (2 fm), its absorption shadows it from encountering subsequent nucleons. The coherence length of the virtual photon’s fluctuation is $`l_{\mathrm{coh}.}1/2m_Nx`$ where $`m_N`$ is the nucleon mass. Therefore the onset of shadowing is expected and observed at $`x0.05`$. In this Gribov multiple scattering picture (?), there is a close relation between shadowing and diffraction. The so-called AGK cutting rules (?) relate the first nuclear shadowing correction to the cross-section for diffractively producing a final state in coherent scattering off a nucleon (integrated over all diffractive final states). See Fig. 13(a) for an illustration of this correspondence. With these relations (and higher order re-scattering generalizations of these) and with the HERA diffractive DIS data as input, the NMC nuclear shadowing data can be reproduced as shown in the sample computation (?, ?) in Fig. 13(b). In the infinite momentum frame (IMF), shadowing arises due to gluon recombination and screening in the target. When the density of partons in the transverse plane of the nucleus becomes very large, many body recombination and screening effects compete against the growth in the cross-section, leading eventually to a saturation of the gluon density (?). In the IMF picture, one can again use the AGK rules we discussed previously to relate shadowing and diffraction (?), and the result is amenable to a partonic interpretation. The saturation regime is characterized by a scale $`Q_s(x,A)`$, called the saturation scale, which grows with decreasing $`x`$ and increasing $`A`$. This saturation scale arises naturally in the Color Glass Condensate (CGC) framework which is discussed in section 3. A natural consequence of saturation physics is the phenomenon of geometrical scaling. (See for instance the discussion on geometrical scaling of HERA data in section 2.1.) It has been argued that the NMC DIS data also display geometrical scaling (?)–the evidence here albeit interesting is not compelling owing to the paucity of nuclear data over a wide range of $`x`$ and $`Q^2`$. It is widely believed that shadowing is a leading twist effect (?, ?), but some of the IMF discussion in the CGC saturation framework suggests higher twist effects are important for $`Q^2Q_s^2`$ because of the large gluon density (?). Constraints from non-linear corrections to the DGLAP framework have also been discussed recently (?). The available data on the $`Q^2`$ dependence of shadowing are inconclusive at small $`x`$. Our empirical definition of shadowing in DIS refers to quark shadowing. Likewise for quarks and anti–quarks in the Drell–Yan process in hadronic collisions. In DIS gluon distributions are inferred only indirectly because the virtual photon couples to quarks. The most precise extractions of gluon distributions thus far are from scaling violations of $`F_2^A`$. To do this properly, one needs a wide window in $`x`$ and $`Q^2`$. In contrast to the highly precise data on nucleon gluon distributions from HERA, our knowledge of nuclear gluon structure functions ($`g_A(x,Q^2)`$) is nearly non-existent. This is especially so relative to our knowledge of quark distributions in nuclei. The most precise data on the modification of gluon distributions in nuclei come from two NMC high precision measurements of the ratio of the scaling violations of the structure functions of Tin (Sn) and Carbon (C). The experiments measure ratios $`f_1=F_2^{\mathrm{Sn}}/F_2^\mathrm{C}`$ and $`f_2=\frac{}{\mathrm{ln}Q^2}f_1`$. The ratio $`r=g_{\mathrm{Sn}}/g_\mathrm{C}`$ can be determined (?) from $`f_1`$ and $`f_2`$ and the scaling violations of $`F_2^{\mathrm{Deuterium}}`$ (with minimal assumptions). The result for $`r`$ is shown in Fig. 14. At small $`x`$, gluon shadowing is observed. The trend suggests that gluon shadowing at small $`x`$ is greater than that of $`F_2`$, even though the error bars are too large for a conclusive statement. At larger $`x`$ of $`0.1<x<0.2`$, one observes anti–shadowing of the gluon distributions. This result from scaling violations can be compared to the ratio of gluon distributions extracted from inclusive $`J/\psi `$ production in DIS. The latter assumes the gluon fusion model of $`J/\psi `$-production. The results for $`r`$ from the latter method are consistent with those from scaling violations. The large experimental uncertainties however leave the extent of anti-shadowing in doubt. Other measurements of scaling violations for the ratio of $`F_2^{\mathrm{Sn}}/F_2^\mathrm{C}`$ showed an increase of the ratio with the increase of $`Q^2`$ consistent with predictions (?, ?). The limited data we have may be interpreted to suggest a provocative picture of nuclear parton densities in the $`x0.1`$$`0.2`$ region, which corresponds to distances of $`1`$$`1.5`$ fm, where medium range and short range inter-nucleon forces are expected to be important. In this region, if the gluon and valence quark fields are enhanced while the sea is somewhat suppressed, as some analyses suggest, gluon-induced interactions between nucleons, as well as valence quark interchanges between nucleons may contribute significantly to nuclear binding (?, ?). Nuclear gluon distributions can also be further constrained by inclusive hadron distributions recently measured by the RHIC experiments (?, ?, ?, ?) in Deuteron-Gold scattering at $`\sqrt{s}=200`$ GeV/nucleon. These RHIC results will be discussed in section 3. ### 2.4 Space-time correlations in QCD The space-time picture of DIS processes strongly depends on the value of Bjorken $`x`$. An analysis of electromagnetic current correlators in DIS reveals that one probes the target wave function at space-time points separated by longitudinal distances $`l_{\mathrm{coh}.}`$ and transverse distances $`1/Q`$. At large $`x`$, ($`x>0.2`$), the virtual photon transforms into a strongly interacting state very close to the active nucleon, typically in the middle of the nucleus. If $`Q^2`$ is large enough as well, the produced partonic state interacts weakly with the medium. At smaller $`x`$ ($`x<0.05`$), the longitudinal length scale $`l_{\mathrm{coh}.}`$ exceeds the nuclear size of the heaviest nuclei. At sufficiently small $`x`$ ($`x<0.005`$ for the heaviest nuclei) DIS processes undergo several spatially separated stages. First, the virtual photon transforms into a quark-gluon wave packet well before the nucleus. Time dilation ensures that interactions amongst partons in the wave packet are frozen over large distances. (These can be several hundred fermis at EIC energies.) The partons in the wave packet interact coherently and instantaneously with the target. At high energies, these interactions are eikonal in nature and do not affect the transverse size of the wave packet. Finally, the fast components of the wave packet transform into a hadronic final state when well past the nucleus. This interval could be as large as $`2\nu /\mu ^2`$ where $`\nu `$ is the energy of the virtual photon and $`\mu 1`$ GeV is a soft hadronic scale. Space-time studies thus far have been limited to semi-exclusive experiments that investigate the phenomenon of color transparency, and more generic inclusive studies of quark propagation through nuclei. Both these studies involved fixed targets. They are briefly summarized below. In pQCD, color singlet objects interact weakly with a single nucleon in the target. Additional interactions are suppressed by inverse powers of $`Q^2`$. This phenomenon is called “color transparency” because the nucleus appears transparent to the color singlet projectile (?, ?, ?). At very high energies, even the interaction of small color singlet projectiles with nuclei can be large. In this kinematic region, the phenomenon is termed “color opacity” (?, ?). The earliest study of color transparency in DIS was a study of coherent $`J/\psi `$ photo-production off nuclei (?). The amplitude of the process at small $`t`$ (momentum transfer squared) is approximately proportional to the nuclear atomic number $`A`$. This indicates that the pair that passes through the nucleus is weakly absorbed. For hadronic projectiles, a similar and approximately linear $`A`$-dependence of the amplitude was observed recently for coherent diffraction of 500 GeV pions into two jets (?), consistent with predictions (?). A number of papers (?, ?, ?) predict that the onset of color transparency at sufficiently large $`Q^2`$ will give for the coherent diffractive production of vector mesons $`{\displaystyle \frac{d\sigma (\gamma _L^{}+AV+A)}{dt}}|_{t=0}A^2.`$ (22) For incoherent diffraction at sufficiently large $`t`$ ($`>0.1`$ GeV<sup>2</sup>), they predict, $`R(Q^2){\displaystyle \frac{d\sigma (\gamma _L^{}+AV+A^{})/dt}{Ad\sigma (\gamma _L^{}+NV+N)/dt}}=1.`$ (23) The first measurements of incoherent diffractive production of vector mesons were performed by the E665 collaboration at Fermilab (?). A significant increase of the nuclear transparency, as reflected in the ratio $`R(Q^2)`$, was observed. The limited luminosity and center-of-mass energy however do not provide a statistically convincing demonstration of color transparency. In addition, the results are complicated by large systematic effects. Measurements of the inclusive hadron distribution for different final states as a function of the virtual photon energy $`\nu `$, its transverse momentum squared $`Q^2`$, the fraction $`z_h`$ of the photon energy carried by the hadron, and the nuclear size $`A`$, provide insight into the propagation of quarks and gluons in nuclear media. In addition to the time and length scales discussed previously, the “formation time” $`\tau _h`$ of a hadron is an additional time scale. It is in principle significantly larger than the production time $`1/Q`$ of a color singlet parton. If the formation time is large, the “pre-hadron” can multiple scatter in the nucleus, thereby broadening its momentum distribution, and also suffer radiative energy loss before hadronization. The QCD prediction for transverse momentum broadening resulting from multiple scattering is (for quarks) given by the expression (?) $`\mathrm{\Delta }p_{}^2={\displaystyle \frac{\alpha _SC_F\pi ^2}{2}}xg(x,Q^2)\rho L0.5\alpha _S\left({\displaystyle \frac{L}{5\mathrm{fm}}}\right)\mathrm{GeV}^2.`$ (24) Here, $`C_F=4/3`$ is the color Casimir of the quark, $`\rho `$ is the nuclear matter density and $`L`$ is the length of matter traversed. The Drell-Yan data in Ref. (?) agree with this expression and with the predicted small size of the effect (empirically, $`\mathrm{\Delta }p_{}^20.12`$ GeV<sup>2</sup> for heavy nuclei). One also observes a large difference in the $`A`$ dependence of the transverse momentum of Drell-Yan di-muons relative to those from $`J/\psi `$ and $`\mathrm{{\rm Y}}`$ production and decay (?). In the former process only the incident quark undergoes strong interactions, whereas in the latter, the produced vector mesons interact strongly as well. However, the size of the effect and the comparable broadening of the $`J/\psi `$ and $`\mathrm{{\rm Y}}`$ (albeit the latter is appreciably smaller than the former) need to be better understood. The $`p_{}`$ imbalance of di-jets in nuclear photo-production suggests a significantly larger $`p_{}`$ broadening effect than in $`J/\psi `$ production (?). This suggests non-universal behavior of $`p_{}`$ broadening effects but may also occur from a contamination of the jets by soft fragments. Parton $`p_{}`$ broadening due to multiple scattering may also be responsible for the anomalous behavior of inclusive hadron production in hadron-nucleus scattering at moderate $`p_{}`$ of a few GeV. In this case, the ratio $`R_{\mathrm{pA}}`$ of inclusive hadron production in hadron-nucleus scattering to the same process on a nucleon, is suppressed at low $`p_{}`$ but exceeds unity between $`12`$ GeV. This “Cronin effect” (?) was discovered in proton-nucleus scattering experiments in the late 70’s. The flavor dependence of the Cronin effect provided an early hint that scattering of projectile partons off gluons dominates over scattering off quarks (?). The Cronin effect will be discussed further in section 3 in light of the recent RHIC experiments on Deuteron-Gold scattering (?, ?, ?, ?). The energy loss of partons due to scattering in nuclear matter is complicated by vacuum induced energy loss in addition to the energy loss due to scattering. One computation suggests that vacuum energy loss is the dominant effect (?). For a quark jet, the medium induced energy loss increases quadratically with the length, $`L`$, and is independent of the energy for $`E\mathrm{}`$. For $`L=5`$ fm, the asymptotic energy loss, $`\mathrm{\Delta }E`$ , is estimated to be less than $`1`$ GeV in a cold nuclear medium (?). This makes it difficult to empirically confirm this remarkable $`L`$-dependence of the energy loss. DIS data are qualitatively consistent with small energy loss (?, ?, ?). The data indicate that the multiplicity of the leading hadrons is moderately reduced (by 10%) for virtual photon energies of the order of $`10`$-$`20`$ GeV for scattering off Nitrogen-14 nuclei. At higher energies, the leading multiplicities gradually become $`A`$-dependent, indicating absorption of the leading partons (?, ?, ?, ?, ?). A pQCD description of partonic energy loss in terms of modified fragmentation functions is claimed to describe HERMES data (?). However, at HERMES energies, and perhaps even at EMC energies, descriptions in terms of hadronic re-scattering and absorption are at least as successful (?, ?). As previously discussed, however, the latter descriptions usually require that the color singlet “pre-hadrons” have a formation time $`\tau _h0.5`$ fm (?, ?, ?). Fig. 15 shows the results from one such model as a function of $`z_h`$ (the fraction of the parton momentum carried by a hadron) and $`\nu `$ (the virtual photon energy) compared to the HERMES data. At EIC energies, a pQCD approach in terms of modified fragmentation functions (?, ?) should be more applicable. The results from these analyses will provide an important test of jet quenching in hot matter descriptions of the RHIC data. ## 3 Scientific Opportunities with an Electron-Ion Collider This section will discuss the exciting scientific opportunities that will be made possible by the novel features of an electron-ion collider: the high luminosity, the possibility to do scans over a wide range in energy, polarization of the electron and hadronic beams, a range of light and heavy nuclear beams and, not least, the collider geometry of the scattering. Scientific firsts for the electron-ion collider will include a) the first high energy polarized electron-polarized proton collider, and b) the first high energy electron-nucleus collider. ### 3.1 Unpolarized ep collisions at EIC Unpolarized e-p collisions have been studied extensively most recently at the HERA collider at DESY. In the eRHIC option for an EIC, the center of mass energy in an e-p collision is anticipated to be $`\sqrt{s}=100`$ GeV compared to $`\sqrt{s}`$ of over 300 GeV at HERA. Although the $`x`$-$`Q^2`$ reach of an EIC may not be as large as that of HERA, it has significant other advantages which we will itemize below. * The current design luminosity is approximately 25 times the design luminosity of HERA. Inclusive observables will be measured with great precision. The additional luminosity will be particularly advantageous for studying semi-inclusive and exclusive final states. * The EIC (particularly in the eRHIC version) will be able to vary the energies of both the electron and nucleon beams. This will enable a first measurement of $`F_L`$ in the small $`x`$ regime. The $`F_L`$ measurement is very important in testing QCD fits of structure functions. * Electron-Deuteron collisions, with tagging of spectator nucleons, will allow high precision studies of the flavor dependence of parton distributions. * An eRHIC detector proposed by Caldwell et al. (?) would have a rapidity coverage nearly twice that of the ZEUS and H1 detectors at HERA. This would allow the reconstruction of the event structure of hard forward jets with and without rapidity gaps in the final state. With this detector, exclusive vector meson and DVCS measurements can be performed for a wider range of the photon-proton center of mass energy squared $`W^2`$. It also permits measurements up to high $`|t|`$, where $`t`$ denotes the square of the difference in four-momenta of the incoming and outgoing proton. These will enable a precise mapping of the energy dependence of final states, as well as open a window into the spatial distribution of partons down to very low impact parameters. We will briefly discuss the physics measurements that can either be done or improved upon with the above enumerated capabilities of EIC/eRHIC. For inclusive measurements, $`F_L`$ is clearly a first, “gold plated” measurement. Current QCD fits predict that $`F_L`$ is very small (and in some analyses negative) at small $`x`$ and small $`Q^2<2`$ GeV<sup>2</sup>. An independent measurement can settle whether this reflects poor extrapolations of data (implying leading twist interpretations of data are still adequate in this regime), or whether higher twist effects are dominant. It will also constrain extractions of the gluon distribution because $`F_L`$ is very sensitive to it. Another novel measurement would be that of structure functions in the region of large $`x1`$. These measurements can be done with 1 fb<sup>-1</sup> of data for up to $`x=0.9`$ and for $`Q^2<250`$ GeV<sup>2</sup>. This kinematic window is completely unexplored to date. These studies can test perturbative QCD predictions of the helicity distribution of the valence partons in a proton (?) as well as the detailed pattern of SU(6) symmetry breaking (?). Moments of structure functions can be compared to lattice data. These should help quantify the influence of higher twist effects. Finally ideas such as Bloom-Gilman duality can be further tested in this kinematic region (?). At small $`x`$, very little is understood about the quark sea. For instance, the origins of the $`\overline{u}\overline{d}`$ asymmetry and the suppression of the strange sea are not clear. High precision measurements of $`\pi ^\pm `$, $`K^\pm `$, $`K_s`$ and open charm will help separate valence and sea contributions in the small-$`x`$ region. We have already discussed Generalized Parton Distributions and DVCS measurements. The high luminosity, wide coverage and measurements at high $`|t|`$ will quantify efforts to extract a 3-D snapshot of the distribution of partons in the proton. ### 3.2 Polarized $`ep`$ collisions at EIC We expect the EIC to dramatically extend our understanding of the spin structure of the proton through measurements of the spin structure function $`g_1`$ over a wide range in $`x`$ and $`Q^2`$, of its parity-violating counterparts $`g_4`$ and $`g_5`$, of gluon polarization $`\mathrm{\Delta }G`$, as well as through spin-dependent semi-inclusive measurements, the study of exclusive reactions, and of polarized photo-production. #### 3.2.1 Inclusive spin-dependent structure functions We have emphasized in Sec. 2.2.1 the need for further measurements of $`g_1(x,Q^2)`$ at lower $`x`$ and higher $`Q^2`$. A particularly important reason is that one would like to reduce the uncertainty in the integral $`\mathrm{\Gamma }_1(Q^2)`$ and hence in $`\mathrm{\Delta }\mathrm{\Sigma }`$. However, the behavior of $`g_1(x,Q^2)`$ at small $`x`$ is by itself of great interest in QCD. At very high energies, Regge theory gives guidance to the expected behavior of $`g_1(x)`$. The prediction (?) is that $`g_1(x)`$ is flat or even slightly vanishing at small $`x`$, $`g_1(x)x^\alpha `$ with $`0.5\alpha 0`$. It is an open question how far one can increase $`Q^2`$ or decrease energy and still trust Regge theory. A behavior of the form $`g_1x^\alpha `$ with $`\alpha <0`$ is unstable under DGLAP evolution (?, ?) in the sense that evolution itself will then govern the small-$`x`$ behavior at higher $`Q^2`$. Under the assumption that Regge theory expectations are realistic at some (low) scale $`Q_0`$ one then obtains “perturbative predictions” for $`g_1(x,Q^2)`$ at $`QQ_0`$. The fixed-target polarized DIS data indicate that although the non-singlet combination $`g_1^pg_1^n`$ is quite singular at small $`x`$, the singlet piece does appear to be rather flat (?, ?), so that the above reasoning applies here. It turns out that the leading eigenvector of small-$`x`$ evolution is such that the polarized quark singlet distribution and gluon density become of opposite sign. For a sizeable positive gluon polarization, this leads to the striking feature that the singlet part of $`g_1(x,Q^2)`$ is negative at small $`x`$ and large $`Q^2`$ (?), driven by $`\mathrm{\Delta }g`$. Figure 16 shows this dramatic behavior for different values of $`Q^2=2,10,20,100`$ GeV<sup>2</sup> (?). The projected statistical uncertainties at eRHIC, corresponding to 400 pb<sup>-1</sup> integrated luminosity with an almost $`4\pi `$ acceptance detector, are also shown. Note that 400 pb<sup>-1</sup> can probably be collected within about one week of e-p running of eRHIC. Thus, at eRHIC one will be well positioned to explore the evolution of the spin structure function $`g_1(x,Q^2)`$ at small $`x`$. We note that at small $`x`$ and toward $`Q^20`$, one could also study the transition region between the Regge and pQCD regimes (?). Predictions for the small-$`x`$ behavior of $`g_1`$ have also been obtained from a perturbative resummation of double-logarithms $`\alpha _s^k\mathrm{ln}^{2k}\left(1/x\right)`$ appearing in the splitting functions (?, ?, ?, ?) at small $`x`$ in perturbative QCD. Some of these calculations indicate a very singular asymptotic behavior of $`g_1(x)`$. It has been shown (?), however, that subleading terms may still be very important even far below $`x=10^3`$. The neutron $`g_1`$ structure function could be measured at eRHIC by colliding the electrons with polarized Deuterons or with Helium. If additionally the hadronic proton fragments are tagged, a very clean and direct measurement could be performed. As can be seen from Fig. 7, information on $`g_1^n`$ at small $`x`$ is scarce. The small-$`x`$ behavior of the isotriplet $`g_1^pg_1^n`$ is particularly interesting for the Bjorken sum rule and because of the steep behavior seen in the fixed-target data (?, ?). It is estimated (?) that an accuracy of the order of $`1\%`$ could be achieved for the Bjorken sum rule in a running time of about one month. One can also turn this argument around and use the accurate measurement of the non-singlet spin structure function and its evolution to determine the value of the strong coupling constant $`\alpha _s(Q^2)`$(?, ?). This has been tried, and the value one gets from this exercise is comparable to the world average for the strong coupling constant. It is expected that if precision low-$`x`$ data from the EIC is available and the above mentioned non-singlet structure functions are measured along with their evolutions, this may result in the most accurate value of the strong coupling constant $`\alpha _s(Q^2)`$. Because of eRHIC’s high energy, very large $`Q^2`$ can be reached. Here, the DIS process proceeds not only via photon exchange; also the $`W`$ and $`Z`$ contribute significantly. Equation 2 shows that in this case new structure functions arising from parity violation contribute to the DIS cross section. These structure functions contain very rich additional information on parton distributions (?, ?, ?). As an example, let us consider charged-current (CC) interactions. Events in the case of $`W`$ exchange are characterized by a large transverse momentum imbalance caused by the inability to detect neutrinos from the event. The charge of the $`W`$ boson is dictated by that of the lepton beam used in the collision. For $`W^{}`$ exchange one then has for the structure functions $`g_1`$ and $`g_5`$ in Equation 6: $$g_1^W^{}(x)=\mathrm{\Delta }u(x)+\mathrm{\Delta }\overline{d}(x)+\mathrm{\Delta }\overline{s}(x),g_5^W^{}(x)=\mathrm{\Delta }u(x)\mathrm{\Delta }\overline{d}(x)\mathrm{\Delta }\overline{s}(x).$$ (25) These appear in the double-spin asymmetry as defined in Ref. (?), where the asymmetry can be expressed in terms of structure functions as $$A^W^{}=\frac{2bg_1^W^{}+ag_5^W^{}}{aF_1^W^{}+bF_3^W^{}}.$$ (26) Here $`a=2(y^22y+2)`$, $`b=y(2y)`$ and $`F_3`$ is the unpolarized parity-violating structure function of Equation 2. Note that the typical scale in the parton densities is $`M_W`$ here. Availability of polarized neutrons and positrons is particularly desirable. For example, one finds at lowest order: $`g_1^{W^{},p}g_1^{W^+,p}`$ $`=`$ $`\mathrm{\Delta }u_v\mathrm{\Delta }d_v`$ (27) $`g_5^{W^+,p}+g_5^{W^{},p}`$ $`=`$ $`\mathrm{\Delta }u_v+\mathrm{\Delta }d_v`$ (28) $`g_5^{W^+,p}g_5^{W^{},n}`$ $`=`$ $`\left[\mathrm{\Delta }u+\mathrm{\Delta }\overline{u}\mathrm{\Delta }d\mathrm{\Delta }\overline{d}\right].`$ (29) The last of these relations gives, after integration over all $`x`$ and taking into account the first-order QCD correction (?), $$_0^1𝑑x\left[g_5^{W^+,p}g_5^{W^{},n}\right]=\left(1\frac{2\alpha _s}{3\pi }\right)g_A,$$ (30) equally fundamental as the Bjorken sum rule. A Monte Carlo study, including the detector effects, has shown that the measurement of the asymmetry in Equation 26 and the parity violating spin structure functions is feasible at eRHIC. Figure 17 shows simulations (?) for the asymmetry and the structure function $`g_5`$ for CC events with an electron beam. The luminosity was assumed to be 2 fb<sup>-1</sup>. The simulated data shown are for $`Q^2>225`$ GeV<sup>2</sup>. Similar estimates exist for $`W^+`$. Measuring this asymmetry would require a positron beam. The curves in the figure use the polarized parton distributions of (?). It was assumed that the unpolarized structure functions will have been measured well by HERA by the time this measurement would be performed at eRHIC. Standard assumptions used by the H1 collaboration about the scattered electrons for good detection were applied. The results shown could be obtained (taking into account machine and detector inefficiencies) in a little over one month with the eRHIC luminosity. It is possible that only one or both of the electron-proton and positron-proton collisions could be performed, depending on which design of the accelerator is finally chosen (see Section 4). #### 3.2.2 Semi-inclusive measurements As we discussed in Sec. 2.2.1, significant insights into the nucleon’s spin and flavor structure can be gained from semi-inclusive scattering $`epehX`$. Knowledge of the identity of the produced hadrons $`h`$ allows separation of the contributions from the different quark flavors. In fixed target experiments, the so-called current hadrons are at forward angles in the laboratory frame. This region is difficult to instrument adequately, especially if the luminosity is increased to gain significant statistical accuracy. A polarized ep collider has the ideal geometry to overcome these shortfalls. The collider kinematics open up the final state into a large solid angle in the laboratory which, using an appropriately designed detector, allows complete identification of the hadronic final state both in the current and target kinematic regions of fragmentation phase space. At eRHIC energies the current and target kinematics are well separated and may be individually studied. At eRHIC higher $`Q^2`$ will be available than in the fixed-target experiments, making the observed spin asymmetries less prone to higher-twist effects, and the interpretation cleaner. Figure 18 shows simulations (?) of the precision with which one could measure the polarized quark and antiquark distributions at the EIC. The events were produced using the DIS generator LEPTO. The plotted uncertainties are statistical only. The simulation was based on an integrated luminosity of 1 fb<sup>-1</sup> for 5 GeV electrons on 50 GeV protons, with both beams polarized to 70$`\%`$. Inclusive and semi-inclusive asymmetries were analyzed using the leading order “purity” method developed by the SMC (?) and Hermes (?) collaborations. Excellent precision for $`\mathrm{\Delta }q/q`$ can be obtained down to $`x0.001`$. The measured average $`Q^2`$ values vary as usual per $`x`$ bin; they are in the range $`Q^2=1.1`$ GeV<sup>2</sup> at the lowest $`x`$ to $`Q^240`$ GeV<sup>2</sup> at high $`x`$. With proton beams, one has greater sensitivity to up quarks than to down quarks. Excellent precision for the down quark polarizations could be obtained by using deuteron or helium beams. With identified kaons, and if the up and down quark distributions are known sufficiently well, one will have a very good possibility to determine the strange quark polarization. As we discussed in Sec. 2.2.1, $`\mathrm{\Delta }s(x)`$ is one of the most interesting quantities in nucleon spin structure. On the right-hand-side of Figure 18, we show results expected for $`\mathrm{\Delta }s(x)`$ as extracted from the spin asymmetries for $`K^\pm `$ production. As in the previous figure, only statistical uncertainties are indicated. The results are compared with the precision available in the Hermes experiment. There is also much interest in QCD in more refined semi-inclusive measurements. For example, the transverse momentum of the observed hadron may be observed. Here, interesting azimuthal-angle dependences arise at leading twist (?, ?), as we discussed in Subsec. 2.2.3. At small transverse momenta, resummations of large Sudakov logarithms are required (?). Measurements at eRHIC would extend previous results from HERA (?) and be a testing ground for detailed studies in perturbative QCD. #### 3.2.3 Measurements of the polarized gluon distribution $`\mathrm{\Delta }g(x,Q^2)`$ One may extract $`\mathrm{\Delta }g`$ from scaling violations of the structure function $`g_1(x,Q^2)`$. Figure 8 shows that indeed some initial information on $`\mathrm{\Delta }g(x,Q^2)`$ has been obtained in this way, albeit with very poor accuracy. The uncertainty of the integral of $`\mathrm{\Delta }g`$ is probably about $`100\%`$ at the moment (?). Measurements at RHIC will vastly improve on this. eRHIC will offer independent and complementary information. Thanks to the large lever arm in $`Q^2`$, and to the low $`x`$ that can be reached, scaling violations alone will constrain $`\mathrm{\Delta }g(x,Q^2)`$ and its integral much better. Studies (?) indicate for example that the total uncertainty on the integral of $`\mathrm{\Delta }G`$ could be reduced to about $`510\%`$ by measurements at eRHIC with integrated luminosity of $`12`$ fb<sup>-1</sup> ($`23`$ years of eRHIC operation). Lepton-nucleon scattering also offers direct ways of accessing gluon polarization. Here one makes use of the photon-gluon fusion (PGF) process, for which the gluon appears at leading order. Charm production is one particularly interesting channel (?, ?, ?). It was also proposed (?, ?, ?) to use jet pairs, produced in the reaction $`\gamma ^{}gq\overline{q}`$, for a determination of $`\mathrm{\Delta }g`$. This process competes with the QCD Compton process, $`\gamma ^{}qqg`$. Feynman diagrams for these processes are shown in Figure 19. In the unpolarized case, dijet production has successfully been used at HERA to constrain the gluon density (?). Dedicated studies have been performed for dijet production in polarized collisions at eRHIC (?), using the MEPJET (?) generator. The two jets were required to have transverse momenta $`>3`$ GeV, pseudorapidities $`3.5\eta 4`$, and invariant mass $`s_{JJ}>100`$ GeV<sup>2</sup>. A $`4\pi `$ detector coverage was assumed. The results for the reconstructed $`\mathrm{\Delta }g(x)`$ are shown in Figure 20, assuming luminosities of 1 fb<sup>-1</sup> (left) and 200 pb<sup>-1</sup> (right). The best probe would be in the region $`0.02x0.1`$; at higher $`x`$, the QCD Compton process becomes dominant. This region is indicated by the shaded areas in the figure. The region $`0.02x0.1`$ is similar to that probed at RHIC. Measurements at eRHIC would thus allow an independent determination of $`\mathrm{\Delta }g`$ in a complementary physics environment. Eventually data for the scaling violations in $`g_1(x,Q^2)`$ and for dijet production in DIS will be analyzed jointly. Such a combined analysis would determine the gluon distribution with yet smaller uncertainties. A first preliminary study for eRHIC (?), following the lines of (?), indeed confirms this. #### 3.2.4 Exploring the partonic structure of polarized photons In the photoproduction limit, when the virtuality of the intermediate photon is small, the $`ep`$ cross-section can be approximated by a product of a photon flux and an interaction cross section of the real photon with the proton. Measurements at HERA in the photoproduction limit have led to a significant improvement in our knowledge of the hadronic structure of the photon. The structure of the photon manifests itself in so-called “resolved” contributions to cross sections. We show this in Figure 21 for the case of photoproduction of hadrons. On the left, the photon participates itself in the hard scattering, through “direct” contributions. On the right, the photon behaves like a hadron. This possibility occurs because of (perturbative) short-time fluctuations of the photon into $`q\overline{q}`$ pairs and gluons, and because of (non-perturbative) fluctuations into vector mesons $`\rho ,\varphi ,\omega `$ with the same quantum numbers (?). The resolved contributions have been firmly established by experiments in $`e^+e^{}`$ annihilation and $`ep`$ scattering (?). A unique application of eRHIC would be to study the parton distributions of polarized quasi-real photons, defined as (?, ?) $$\mathrm{\Delta }f^\gamma (x)f_+^{\gamma _+}(x)f_{}^{\gamma _+}(x),$$ (31) where $`f_+^{\gamma _+}`$ $`(f_{}^{\gamma _+})`$ denotes the density of a parton $`f=u,d,s,\mathrm{},g`$ with positive (negative) helicity in a photon with positive helicity. The $`\mathrm{\Delta }f^\gamma (x)`$ give information on the spin structure of the photon; they are completely unmeasured so far. Figure 22 shows samples from studies (?, ?) for observables at eRHIC that would give information on the $`\mathrm{\Delta }f^\gamma (x)`$. Two models for the $`\mathrm{\Delta }f^\gamma (x)`$ were considered (?), one with a strong polarization of partons in the photon (“maximal” set), the other with practically unpolarized partons (“minimal” set). On the left, we show the double-spin asymmetry for photoproduction of high-$`p_T`$ pions, as a function of the pion’s pseudorapidity $`\eta _{\mathrm{lab}}`$ in the eRHIC laboratory frame. The advantage of this observable is that for negative $`\eta _{\mathrm{lab}}`$, in the proton backward region, the photon mostly interacts “directly”, via the process $`\gamma gq\overline{q}`$, whereas its partonic content becomes visible at positive $`\eta _{\mathrm{lab}}`$. This may be seen from the figure, for which we have also used two different sets of polarized parton distributions of the proton (?), mainly differing in $`\mathrm{\Delta }g(x)`$. The right part of Figure 22 shows predictions for the spin asymmetry in dijet photoproduction at eRHIC. If one assumes the jets to be produced by a $`22`$ partonic hard scattering, the jet observables determine the momentum fractions $`x_{p,\gamma }`$ of the partons in the proton and the photon. Selecting events with $`x_\gamma <1`$, one therefore directly extracts the “resolved”-photon contribution. At higher orders, this picture is somewhat diluted, but remains qualitatively intact. Such measurements of dijet photoproduction cross sections at HERA (?) have been particularly successful in providing information on photon structure. This makes the spin asymmetry a good candidate for learning about the $`\mathrm{\Delta }f^\gamma `$ at eRHIC. In the figure we show results for the asymmetry in three different bins of $`x_\gamma `$. One can see that with 1 fb<sup>-1</sup> luminosity one should be able at eRHIC to establish the existence of polarized resolved-photon contributions, and distinguish between our “maximal” and “minimal” photon scenarios. For a first exploration one could also use the approach of “effective” parton densities considered in (?, ?, ?, ?). We finally note that measurements of the polarized total photoproduction cross section at high energies would also give new valuable information on the high-energy contribution to the Drell-Hearn-Gerasimov sum rule (?). The latter relates the total cross sections with photon-proton angular momentum $`3/2`$ and $`1/2`$ to the anomalous magnetic moments of the nucleon (?): $$_0^{\mathrm{}}\frac{d\nu }{\nu }\left[\sigma _{3/2}(\nu )\sigma _{1/2}(\nu )\right]=\frac{2\pi ^2\alpha }{M^2}\kappa ^2=\{\begin{array}{cc}204.5\mu \text{b}\hfill & p\hfill \\ 232.8\mu \text{b}\hfill & n\hfill \end{array},$$ (32) where on the right we have given the numerical values of the sum rule. Currently, the experimental result for the proton is a few percent high, and the one for the neutron about $`20\%`$ low (?). There is practically no information on the contribution to the sum rule from photon energies $`\nu 3`$ GeV; estimates based on Regge theory indicate that it is possible that a substantial part comes from this region. Measurements at eRHIC could give definitive answers here. The H1 and ZEUS detectors at DESY routinely take data using electron taggers situated in the beam pipe 6 - 44 meters away from the end of the detectors. They detect the scattered electrons from events at very low $`Q^2`$ and scattering angles. If electron taggers were included in eRHIC, similar measurements could be performed. The $`Q^2`$ range of such measurements at eRHIC is estimated to be $`10^810^2`$ GeV<sup>2</sup>. #### 3.2.5 Hard exclusive processes As we have discussed in Subsection 2.2.2, generalized parton distributions (GPDs) are fundamental elements of nucleon structure. They contain both the parton distributions and the nucleon form factors as limiting cases, and they provide information on the spatial distribution of partons in the transverse plane. GPDs allow the description of exclusive processes at large $`Q^2`$, among them DVCS. It is hoped that eventually these reactions will provide information on the total angular momenta carried by partons in the proton. See for example Eq. (15). The experimental requirements for a complete investigation of GPDs are formidable. Many different processes need to be investigated at very high luminosities, at large enough $`Q^2`$, with polarization, and with suitable resolution to determine reliably the hadronic final state. The main difficulty, however, for experimental measurements of exclusive reactions is detecting the scattered proton. If the proton is not detected, a “missing-mass” analysis has to be performed. In case of the DVCS reaction, there may be a significant contribution from the Bethe-Heitler process. The amplitude for the Bethe-Heitler process is known and, as we discussed in Subsection 2.2.2, one may construct beam-spin and charge asymmetries to partly eliminate the Bethe-Heitler contribution. Early detector design studies have been performed for the EIC (?). These studies indicate that the acceptance can be significantly increased by adding stations of Silicon-strip-based Roman Pot Detectors away from the central detector in a HERA-like configuration. The detector recently proposed for low-$`x`$ and low-$`Q^2`$ studies at the EIC (?) (for details, see Section 5) may also be of significant use to measure the scattered proton. Further studies are underway and will proceed along with iterations of the design of the interaction region and of the beam line. Although more detailed studies need yet to be performed, we anticipate that the EIC would provide excellent possibilities for studying GPDs. Measurements at the collider will complement those now underway at fixed target experiments and planned with the 12-GeV upgrade at the Jefferson Laboratory (?). ### 3.3 Exploring the nucleus with an electron-ion collider In this section, we discuss the scientific opportunities available with the EIC in DIS off nuclei. At very high energies, the correct degrees of freedom to describe the structure of nuclei are quarks and gluons. The current understanding of partonic structure is just sufficient to suggest that their behavior is non-trivial. The situation is reminiscent of Quantum Electrodynamics. The rich science of condensed matter physics took a long time to develop even though the nature of the interaction was well understood. Very little is known about the condensed matter , many-body properties of QCD, particularly at high energies. There are sound reasons based in QCD to believe that partons exhibit remarkable collective phenomena at high energies. Because the EIC will be the first electron-ion collider, we will be entering a terra incognita in our understanding of the properties of quarks and gluons in nuclei. The range in $`x`$ and $`Q^2`$ and the luminosity will be greater than at any previous fixed target DIS experiment. Further, the collider environment is ideal for studying semi-inclusive and exclusive processes. Finally, it is expected that a wide range of particle species and beam energies will be available to study carefully the systematic variation of a wide range of observables with target size and energy. We will begin our discussion in this section by discussing inclusive ”bread and butter” observables such as the inclusive nuclear quark and gluon structure functions. As we observed previously, very little is known about nuclear structure functions at small $`x`$ and $`Q^2\mathrm{\Lambda }_{\mathrm{QCD}}^20.04`$ GeV<sup>2</sup>. This is especially true of the nuclear gluon distribution. We will discuss the very significant contributions that the EIC can make in rectifying this situation. A first will be a reliable extraction of the longitudinal structure function at small $`x`$. Much progress has been made recently in defining universal diffractive structure functions (?, ?, ?). These structure functions can be measured in nuclei for the first time. Generalized parton distributions will help provide a three-dimensional snapshot of the distribution of partons in the nucleus (?). We will discuss the properties of partons in a nuclear medium and the experimental observables that will enable us to tease out their properties. These include nuclear fragmentation functions that contain valuable information on hadronization in a nuclear environment. The momentum distributions of hadronic final states as functions of $`x`$, $`Q^2`$, and the fraction of the parton energy carried by a hadron also provide insight into dynamical effects such as parton energy loss in the nuclear medium. A consequence of small $`x`$ evolution in QCD is the phenomenon of parton saturation (?). This arises from the competition between attractive Bremsstrahlung (?) and repulsive screening and recombination (many body) effects (?), which results in a phase space density of partons of order $`1/\alpha _S`$. At such high parton densities, the partons in the wavefunction form a Color Glass Condensate (CGC) for reasons we will discuss later (?). The CGC is an effective theory describing the remarkable universal properties of partons at high energies. It provides an organizing principle for thinking about high energy scattering and has important ramifications for colliders. The evolution of multi-parton correlations predicted by the CGC can be studied with high precision in lepton-nucleus collisions. Experimental observables measured at the EIC can be compared and contrasted with observables extracted in proton/Deuteron-nucleus and nucleus-nucleus scattering experiments at RHIC and LHC. The kinematic reach of the EIC will significantly overlap with these experiments. Measurements of parton structure functions and multi-parton correlations in the nuclear wave function will provide a deeper understanding of the initial conditions for the formation of a quark gluon plasma (QGP). Final state interactions in heavy ion collisions such as the energy loss of leading hadrons in hot matter (often termed ”jet quenching”) are considered strong indicators of the formation of the QGP. The EIC will provide benchmark results for cold nuclear matter which will help quantify energy loss in hot matter. Finally, recent results on inclusive hadron production in RHIC D-Au collisions at 200 GeV/nucleon show hints of the high parton density effects predicted by the CGC. We will discuss these and consider the similarities and differences between a p/D-A and an e-A collider. #### 3.3.1 Nuclear parton distributions The range of the EIC in $`x`$ and $`Q^2`$ was discussed previously (see Fig. 2). It is significantly larger than for the previous fixed target experiments. The projected statistical accuracy, per inverse picobarn of data, of a measurement of the ratio $`\frac{R}{\mathrm{ln}Q^2}`$ versus $`x`$ at the EIC relative to data from previous NMC measurements and a hypothetical future e-A collider at HERA energies is shown in Fig. 23. Here $`R`$ denotes the ratio of nuclear structure functions, $`R=F_2^A/F_2^N`$. As discussed previously, the logarithmic derivative with $`Q^2`$ of this ratio can be used to extract the nuclear gluon distribution. The EIC is projected to have an integrated luminosity of several hundred pb<sup>-1</sup> for large nuclei, so one can anticipate high precision measurements of nuclear structure functions at small $`x`$. In particular, because the energy of the colliding beams can be varied, the nuclear longitudinal structure function can be measured for the first time at small $`x`$. At small $`x`$ and large $`Q^2`$, it is directly proportional to the gluon distribution. At smaller values of $`Q^2`$, it may be more sensitive to higher twist effects than $`F_2`$ (?). Measurements of nuclear structure functions in the low $`x`$ kinematic region will test the predictions of the QCD evolution equations in this kinematic region. The results of QCD evolution with $`Q^2`$ depend on input from the structure functions at smaller values of $`Q^2`$ for a range of $`x`$ values. The data on these is scarce for nuclei. These results are therefore very sensitive to models of the small $`x`$ behavior of structure functions at low $`Q^2`$. A nice plot from Ref. (?) reproduced in Fig. 24 clearly illustrates the problem. Figure 24 shows results from theoretical models for the ratio of the gluon distribution in Lead to that in a proton as a function of $`x`$. Though all the models employ the same QCD evolution equations, the range in uncertainty is rather large at small $`x`$ – about a factor of 3 at $`x10^4`$. Although one can try to construct better models, the definitive constraint can only come from experiment. The shadowing of gluon distributions shown in Fig. 24 is not understood in a fundamental way. We list here some relevant questions which can be addressed by a future electron-ion collider. * Is shadowing a leading twist effect, namely, is it unsuppressed by a power of $`Q^2`$ ? Most models of nuclear structure functions at small $`x`$ assume this is the case. (For a review, see Ref. (?).) Is there a regime of $`x`$ and $`Q^2`$, where power corrections due to high parton density effects can be seen? (?, ?, ?) * What is the relation of shadowing to parton saturation? As we will discuss, parton saturation dynamically gives rise to a semi-hard scale in nuclei. This suggests that shadowing at small $`x`$ can be understood in a weak coupling analysis. * Is there a minimum to the shadowing ratio for fixed $`Q^2`$ and $`A`$ with decreasing $`x`$? If so, is it reached faster for gluons or for quarks? * The Gribov relation between shadowing and diffraction that we discussed previously is well established at low parton densities. How is it modified at high parton densities? The EIC can test this relation directly by measuring diffractive structure functions in ep (and e-A) and shadowing in e-A collisions. * Is shadowing universal? For instance, would gluon parton distribution functions extracted from p-A collisions at RHIC be identical to those extracted from e-A in the same kinematic regime? The naive assumption that this is the case may be false if higher twist effects are important. Later in this review, we will discuss the implications of the possible lack of universality for p-A and A-A collisions at the LHC. We now turn to a discussion of diffractive structure functions. At HERA, hard diffractive events were observed where the proton remained intact and the virtual photon fragmented into a hard final state producing a large rapidity gap between the projectile and target. A rapidity gap is a region in rapidity essentially devoid of particles. In pQCD, the probability of a gap is exponentially suppressed as a function of the gap size. At HERA though, gaps of several units in rapidity are relatively unsuppressed; one finds that roughly 10% of the cross–section corresponds to hard diffractive events with invariant masses $`M_X>3`$ GeV. The remarkable nature of this result is transparent in the proton rest frame: a $`50`$ TeV electron slams into the proton and, 10% of the time, the proton is unaffected, even though the interaction causes the virtual photon to fragment into a hard final state. The interesting question in diffraction is the nature of the color singlet object (the “Pomeron”) within the proton that interacts with the virtual photon. This interaction probes, in a novel fashion, the nature of confining interactions within hadrons. (We will discuss later the possibility that one can study in diffractive events the interplay between strong fields produced by confining interactions and those generated by high parton densities.) In hard diffraction, because the invariant mass of the final state is large, one can reasonably ask questions about the quark and gluon content of the Pomeron. A diffractive structure function $`F_{2,A}^{D(4)}`$ can be defined (?, ?, ?), in a fashion analogous to $`F_2`$, as $`{\displaystyle \frac{d^4\sigma _{eAeXA}}{dx_{Bj}dQ^2dx_pdt}}`$ $`=A{\displaystyle \frac{4\pi \alpha _{\mathrm{em}}^2}{xQ^4}}\left\{1y+{\displaystyle \frac{y^2}{2[1+R_A^{D(4)}(\beta ,Q^2,x_p,t)]}}\right\}`$ (33) $`\times `$ $`F_{2,A}^{D(4)}(\beta ,Q^2,x_p,t),`$ where, $`y=Q^2/sx_{Bj}`$, and analogously to $`F_2`$, one has $`R_A^{D(4)}=F_L^{D(4)}/F_T^{D(4)}`$. Further, $`Q^2=q^2>0`$, $`x_{Bj}=Q^2/2Pq`$, $`x_p=q(PP^{})/qP`$, $`t=(PP^{})^2`$ and $`\beta =x_{Bj}/x_p`$. Here $`P`$ is the initial nuclear momentum, and $`P^{}`$ is the net momentum of the fragments $`Y`$ in the proton fragmentation region. Similarly, $`M_X`$ is the invariant mass of the fragments $`X`$ in the electron fragmentation region. An illustration of the hard diffractive event is shown in Fig. 25. It is more convenient in practice to measure the structure function $`F_{2,A}^{D(3)}=F_{2,A}^{D(4)}𝑑t`$, where $`|t_{min}|<|t|<|t_{max}|`$, where $`|t_{min}|`$ is the minimal momentum transfer to the nucleus, and $`|t_{max}|`$ is the maximal momentum transfer to the nucleus that still ensures that the particles in the nuclear fragmentation region $`Y`$ are undetected. An interesting quantity to measure is the ratio $`R_{A1,A2}(\beta ,Q^2,x_p)=\frac{F_{2,A1}^{D(3)}(\beta ,Q^2,x_p)}{F_{2,A2}^{D(3)}(\beta ,Q^2,x_p)}`$. The $`A`$-dependence of this quantity will contain very useful information about the universality of the structure of the Pomeron. In a study for e-A collisions at HERA, it was argued that this ratio could be measured with high systematic and statistical accuracy (?)–the situation for eRHIC should be at least comparable, if not better. Unlike $`F_2`$ however, $`F_2^D`$ is not truly universal–it cannot be applied, for instance, to predict diffractive cross sections in p-A scattering; it can be applied only in other lepton–nucleus scattering studies (?, ?). This has been confirmed by a study where diffractive structure functions measured at HERA were used as an input in computations for hard diffraction at Fermilab. The computations vastly overpredicted the Fermilab data on hard diffraction (?). Some of the topics discussed here will be revisited in our discussion of high parton densities. #### 3.3.2 Space-time evolution of partons in a nuclear environment The nuclear structure functions are inclusive observables and are a measure of the properties of the nuclear wavefunction. Less inclusive observables, which measure these properties in greater detail, will be discussed in the section on the Color Glass Condensate. In addition to studying the wave function, we are interested in the properties of partons as they interact with the nuclear medium. These are often called final state interactions to distinguish them from the initial state interactions in the wavefunction. Separating which effects arise from the wavefunction is not easy because our interpretation of initial state and some final state interactions may depend on the gauge in which the computations are performed (?). Isolating the two effects in experiments is difficult. A case in point is the study of energy loss effects on final states in p-A collisions (?). These effects are not easy to distinguish from shadowing effects in the wavefunction. Nevertheless, in the right kinematics this can be done. In section 2, we discussed various final state in-medium QCD processes such as color transparency, parton energy loss and the medium modification of fragmentation functions. The EIC will enable qualitative progress in studies of the space-time picture of strong interactions relative to previous fixed target DIS experiments. The reasons for this are as follows. * The high luminosity of the EIC will increase by many orders of magnitude the current data sample of final states in DIS scattering off nuclei at high energies. * The EIC will provide a much broader range of $`Q^2`$ and $`x`$, making it possible to compare dynamics for approximately the same space-time coherence lengths as a function of $`Q^2`$. Fixing the coherence length of partons will allow one to distinguish events wherein a photon is transformed into a strongly interacting system either outside or inside the nucleus. This will help isolate initial state interactions from those in the final state. * The collider geometry will enable measurements of final states currently impossible in fixed target kinematics. In particular, a hermetic detector would clearly isolate coherent processes as well as quasi-elastic processes in DIS off nuclei. In addition, one can study the sizes and distributions of rapidity gaps as a function of nuclear size and energy. These will provide a sensitive probe of the interplay between space-time correlations in the final state and in the nuclear wavefunction. * The detection of nucleons produced in the nuclear fragmentation region would make it feasible to study DIS as a function of the number of the nucleons involved in the interaction. In particular, it may be possible to study impact parameter dependence of final states, which will be important to understand in detail the nuclear amplification of final state effects. In addition, the impact parameter dependence will help distinguish geometrical effects from dynamical effects in event-by-event studies of final states. In section 2, we discussed Generalized Parton Distributions (GPDs) in the context of DIS scattering off nucleons. These GPDs can also be measured in DIS scattering off nuclei (?). The simplest system in which to study GPDs is the deuteron. The transition from $`Dp+n`$ in the kinematics where the neutron absorbs the momentum transfer in the scattering is sensitive to the GPD in the neutron with the proton playing the role of a spectator (?). Some preliminary studies have been done for heavier nuclei (?). Certain higher twist correlations in nuclei which scale as $`A^{4/3}`$ are sensitive to nucleon GPDs (?). This leads us to a discussion of GPDs in nuclei at small $`x`$. As we will discuss, high parton density effects are enhanced in large nuclei. $`k_{}`$ dependent GPDs might provide the right approach to study this novel regime (?). The study of nuclear GPDs at moderate and small $`x`$ is a very promising, albeit nascent, direction for further research to uncover the detailed structure of hard space-time processes in nuclear media. These nuclear distributions can be studied for the first time with the EIC. #### 3.3.3 The Color Glass Condensate The Color Glass Condensate (CGC) is an effective field theory describing the properties of the dominant parton configurations in hadrons and nuclei at high energies (?). The degrees of freedom are partons–which carry color charge–hence the “Color” in CGC. The matter behaves like a glass for the following reason. The kinematics of high energy scattering dictates a natural separation between large $`x`$ and small $`x`$ modes (?). The large $`x`$ partons at high energies behave like frozen random light cone sources over time scales that are large compared to the dynamical time scales associated with the small-$`x`$ partons. One can therefore describe an effective theory where the small $`x`$ partons are dynamical fields and the large-$`x`$ partons are frozen sources (?). Under quantum evolution (?), this induces a stochastic coupling between the wee partons via their interaction with the sources. This stochastic behavior is very similar to that of a spin glass. Finally, the Condensate in CGC arises because each of these colored configurations is very similar to a Bose-Einstein Condensate. The occupation number of the gluons can be computed to be of order $`1/\alpha _S`$, and the typical momentum of the partons in the configuration is peaked about a typical momentum–the saturation momentum $`Q_s`$. These properties are further enhanced by quantum evolution in $`x`$. Because the occupation number is so large, by the correspondence principle of quantum mechanics, the small $`x`$ modes can be treated as classical fields. The classical field retains its structure while the saturation scale, generated dynamically in the theory, grows with energy: $`Q_s(x^{})>Q_s(x)`$ for $`x^{}<x`$. The CGC is sometimes used interchangeably with “saturation” (?) – both refer to the same phenomenon, the behavior of partons at large occupation numbers. The Jalilian-Marian-Iancu-McLerran-Weigert-Leonidov-Kovner (JIMWLK) renormalization group equations describe the properties of partons in the high density regime (?). They form an infinite hierarchy (analogous to the Bogoliubov-Born-Green-Kirkwood-Young (BBGKY) hierarchy in statistical mechanics) of ordinary differential equations for the gluon correlators $`A_1A_2\mathrm{}A_n_Y`$, where $`Y=\mathrm{ln}(1/x)`$ is the rapidity. Thus the evolution, with $`x`$, of multi-gluon (semi-inclusive) final states provides precise tests of these equations. The full hierarchy of equations are difficult to solve<sup>2</sup><sup>2</sup>2For a preliminary numerical attempt, see Ref. (?). though there have been major theoretical developments in that direction recently (?). A mean field version of the JIMWLK equation, called the Balitsky-Kovchegov (BK) equation (?), describes the inclusive scattering of the quark-anti-quark dipole off the hadron in deeply inelastic scattering. In particular, the virtual photon-proton cross-section at small $`x`$ can be written as (?, ?) $$\sigma _{T,L}^{\gamma ^{}p}=d^2r_{}𝑑z|\psi _{T,L}(r_{},z,Q^2)|^2\sigma _{q\overline{q}N}(r_{},x),$$ (34) where $`|\psi _{T,L}|^2`$ is the probability for a longitudinally (L) or transversely (T) polarized virtual photon to split into a quark with momentum fraction $`z`$ and an anti-quark with momentum fraction $`1z`$ of the longitudinal momentum of the virtual photon. For the quark and anti-quark located at $`\stackrel{}{x}_{}`$ and $`\stackrel{}{y}_{}`$ respectively from the target, their transverse size is $`\stackrel{}{r}_{}=\stackrel{}{x}_{}\stackrel{}{y}_{}`$, and the impact parameter of the collision is $`\stackrel{}{b}=(\stackrel{}{x}_{}+\stackrel{}{y}_{})/2`$. The probability for this splitting is known exactly from QED and it is convoluted with the cross-section for the $`q\overline{q}`$-pair to scatter off the proton. This cross-section for a dipole scattering off a target can be expressed as $$\sigma _{q\overline{q}N}(x,r_{})=2d^2b𝒩_Y(x,r_{},b),$$ (35) where $`𝒩_Y`$ is the imaginary part of the forward scattering amplitude. The BK equation (?) for this amplitude has the operator form $$\frac{𝒩_Y}{Y}=\overline{\alpha }_S𝒦_{\mathrm{BFKL}}\left\{𝒩_Y𝒩_Y^2\right\}.$$ (36) Here $`𝒦_{\mathrm{BFKL}}`$ is the well known Balitsky-Fadin-Kuraev-Lipatov (BFKL) kernel (?). When $`𝒩<<1`$, the quadratic term is negligible and one has BFKL growth of the number of dipoles; when $`𝒩`$ is close to unity, the growth saturates. The approach to unity can be computed analytically (?). The BK equation is the simplest equation including both the Bremsstrahlung responsible for the rapid growth of amplitudes at small $`x`$ as well as the repulsive many body effects that lead to a saturation of this growth. Saturation models, which incorporate key features of the CGC, explain several features of the HERA data. In section 2.1, we discussed the property of geometrical scaling observed at HERA which is satisfied by the LHS of Eq. 34, where it scaled as a function of the ratio of $`Q^2`$ to the saturation scale $`Q_s^2`$. We also mentioned briefly a simple saturation model, the Golec-Biernat model (?), which captured essential features of this phenomenon in both inclusive and diffractive cross sections at HERA. Geometric scaling arises naturally in the Color Glass Condensate (?, ?), and it has been studied extensively both analytically (?) and numerically (?, ?, ?) for the BK equation. The success of saturation models, as discussed in section 2.1, in explaining less inclusive features of the HERA data is also encouraging since their essential features can be understood to follow from the BK equation. Below we will discuss the implications of mean field studies with the BK equation, as well as effects beyond BK. As mentioned previously, a very important feature of saturation is the dynamical generation of a dimensionful scale $`Q_s^2\mathrm{\Lambda }_{\mathrm{QCD}}^2`$, which controls the running of the coupling at high energies: $`\alpha _S(Q_s^2)1`$. From the BK equation, or more generally, from solutions of BFKL in the presence of an absorptive boundary (corresponding to a CGC-like regime of high parton densities), one can deduce that, for fixed coupling, $`Q_s^2`$ has the asymptotic form $`Q_s^2=Q_0^2\mathrm{exp}\left(cY\right)`$, where $`c=4.8\alpha _S`$ and $`Y=\mathrm{ln}(x_0/x)`$. Here, $`Q_0^2`$ and $`x_0`$ are parameters from the initial conditions. Pre-asymptotic $`Y`$ dependent corrections can also be computed and are large. The behavior of $`Q_s^2`$ changes qualitatively when running coupling effects are taken into account. The state of the art is a computation of the saturation scale to next-to-leading order in BFKL with additional resummation of collinear terms that stabilize the predictions of NLO BFKL (?). One recovers the form $`Q_s^2=Q_0^2\mathrm{exp}\left(\lambda Y\right)`$, now with small pre-asymptotic corrections, with $`\lambda 0.25`$. Remarkably, this value is very close to the value extracted in the Golec-Biernat model from fits to the HERA data. Fig. 26 shows a schematic plot of the CGC and extended scaling regions in the $`x`$-$`Q^2`$ plane. Clearly, with the wide kinematic range of the EIC, and the large number of available measurements–to be discussed later–one has the opportunity to make this plot quantitative. One can further add an additional axis for the atomic number to see how the kinematic reach of the CGC scales with $`A`$. In principle, one can also study the impact parameter dependence of the saturation scale in addition to the $`A`$-dependence. #### 3.3.4 Signatures of the CGC Inclusive signatures. Inclusive measurements include $`F_2`$ and $`F_L`$ for a wide range of nuclei, the latter measurements being done independently for the first time. The data will be precise enough to extract derivatives of these with respect to $`\mathrm{ln}Q^2`$ and $`\mathrm{ln}x`$ in a wide kinematic range in $`x`$ and $`Q^2`$. Logarithmic derivatives of $`F_2`$ and $`F_L`$ will enable the extraction of the coefficient $`\lambda `$ of the saturation scale, which as discussed previously, is defined to be $`Q_s^2=Q_0^2e^{\lambda Y}`$, where $`Y=\mathrm{ln}(x_0/x)`$, and where $`x_0`$ and $`Q_0^2`$ are reference values corresponding to the initial conditions for small-$`x`$ evolution. Simulations suggest that a precise extraction of this quantity may be feasible (?). Except at asymptotic energies, $`\lambda \lambda (Y)`$. Predictions exist for “universal” pre-asymptotic $`Y`$-dependent corrections to $`\lambda `$ (?). Second derivatives of $`F_2`$ and $`F_L`$ with respect to $`\mathrm{ln}(x_0/x)`$ will be sensitive to these corrections. The logarithmic derivatives of $`F_2`$ and $`F_L`$ with $`Q^2`$, especially the latter, will be sensitive to higher twist effects for $`Q^2Q_s^2(x,A)`$. The saturation scale is larger for smaller $`x`$ and larger $`A`$–thus deviations of predictions of CGC fits from DGLAP fits should systematically increase as a function of both. CGC fits have been shown to fit HERA data at small $`x`$ (?, ?). These fits can be extended to nuclei and compared to scaling violation data relative to DGLAP fits. The $`A`$ dependence of the saturation scale can also be extracted from nuclear structure functions at small $`x`$. Again, predictions exist for the pre-asymptotic scaling of the saturation scale with rapidity (or $`x`$), for different $`A`$ (?), that can be tested against the data. In the BK equation (mean field approximation of the CGC renormalization group equations), we now have a simple way to make predictions for the effects of high parton densities on both inclusive and diffractive (?) structure functions. There are now a few preliminary computations for e-A DIS in this framework (?, ?). Much more remains to be done–in particular, comparisons with DGLAP for EIC kinematics and detector cuts. Semi-inclusive and exclusive signatures. The collider geometry of the EIC will greatly enhance the semi-inclusive final states in e-A relative to previous fixed target experiments. Inclusive hadron production at $`p_{}Q_s`$ should be sensitive to higher twist effects for $`Q^2Q_s^2(x,A)`$. For the largest nuclei, these effects should be clearly distinguishable from DGLAP based models. Important semi-inclusive observables are coherent (or diffractive) and inclusive vector meson production, which are sensitive measures of the nuclear gluon density (?, ?). Exclusive vector meson production was suggested by Mueller, Munier and Stasto (?) as a way to extract the S-matrix (and therefore the saturation scale in the Golec-Biernat–Wüsthoff parameterization) from the $`t`$-dependence of exclusive $`\rho `$-meson production. A similar analysis of $`J/\psi `$ production was performed by Guzey et al. (?). These studies for e-A collisions will provide an independent measure of the energy dependence of the saturation scale in nuclei. An extensive recent theoretical review of vector meson production of HERA (relevant for EIC studies as well) can be found in Ref. (?). In hard diffraction, for instance, one should be able to distinguish predictions based on the strong field effects of BK (or hard pomeron based approaches in general) from the soft pomeron physics associated with confinement. As we discussed previously, some saturation models predict that hard diffractive events will constitute 30-40% of the cross-section (?, ?). These computations can be compared with DGLAP predictions which match soft Pomeron physics with hard perturbative physics. One anticipates that the latter would result in a much smaller fraction of the cross-section and should therefore be easily distinguishable from CGC based “strong field” diffraction. The BK renormalization group equation is not sensitive to multi-particle correlations. These are sensitive to effects such as Pomeron loops (?), although phenomenological consequences of these remain to be explored. These effects are reflected in multiplicity fluctuations and rapidity correlations over several units in rapidity (?, ?). One anticipates quantitative studies of these will be developed in the near future. A wide detector coverage able to resolve the detailed structure of events will be optimal for extracting signatures of the novel physics of high parton densities. #### 3.3.5 Exploring the CGC in proton/Deuteron-nucleus collisions Although high parton density hot spots may be studied in pp collisions, they are notoriously hard to observe. The proton is a dilute object, except at small impact parameters, and one needs to tag on final states over a wide $`4\pi `$ coverage. Deuteron-nucleus experiments are more promising in this regard. They have been performed at RHIC and may be performed at LHC in the future. The Cronin effect discovered in the late 70’s (?) predicts a hardening of the transverse momentum spectrum in proton-nucleus collisions, relative to proton-proton collisions at transverse momenta of order $`p_{}12`$ GeV. It disappears at much larger $`p_{}`$. A corresponding depletion is seen at low transverse momenta. The effect was interpreted as arising from the multiple scatterings of partons from the proton off partons from the nucleus (?). First data from RHIC on forward D-Au scattering at $`\sqrt{s}=200`$ GeV/nucleon demonstrate how the Cronin effect is modified with energy or, equivalently, with the rapidity. The $`x`$ values in nuclei probed in these experiments, at $`p_{}2`$ GeV, range from $`10^2`$ in the central rapidity region down to $`10^4`$ at very forward rapidities<sup>3</sup><sup>3</sup>3It has been argued (?), however, that the forward D-A cross section in the BRAHMS kinematic regime receives sizable contributions also from rather large $`x`$ values. At central rapidities, one clearly sees a Cronin peak at $`p_{}12`$ GeV. A dramatic result obtained by the BRAHMS (?) experiment at RHIC<sup>4</sup><sup>4</sup>4The trends seen by BRAHMS are also well corroborated by the PHOBOS, PHENIX and STAR experiments at RHIC in different kinematic ranges (?, ?, ?). is the rapid shrinking of the Cronin peak with rapidity shown in Fig. 27. In Fig. 28, the centrality dependence of the effect is shown. At central rapidities, the Cronin peak is enhanced in more central collisions. For forward rapidities, the trend is reversed: more central collisions at forward rapidities show a greater suppression than less central collisions! Parton distributions in the classical theory of the CGC exhibit the Cronin effect (?, ?, ?). However, unlike this classical Glauber picture (?), quantum evolution in the CGC shows that it breaks down completely when the $`x_2`$ in the target is such that $`\mathrm{ln}(1/x_2)1/\alpha _S`$. This is precisely the trend observed in the RHIC D-Au experiments (?). The rapid depletion of the Cronin effect in the CGC picture is due to the onset of BFKL evolution, whereas the subsequent saturation of this trend reflects the onset of saturation effects (?). The inversion of the centrality dependence can be explained as arising from the onset of BFKL anomalous dimensions, that is, the nuclear Bremsstrahlung spectrum changes from $`Q_s^2/p_{}^2Q_s/p_{}`$. Finally, an additional piece of evidence in support of the CGC picture is the broadening of azimuthal correlations (?) for which preliminary data now exists from the STAR collaboration (?). We note that alternative explanations have been given to explain the BRAHMS data (?). These ideas can be tested conclusively in photon and di-lepton production in D-A collisions at RHIC (?) as well as by more detailed correlation studies. Hadronic collisions in pQCD are often interpreted within the framework of collinear factorization. At high energies, $`k_{}`$ factorization may be applicable (?) where the relevant quantities are “unintegrated” $`k_{}`$ dependent parton densities. Strict $`k_{}`$-factorization which holds for gluon production in p-A collisions (?, ?) is broken for quark production (?, ?), for azimuthal correlations (?) and diffractive final states (?). For a review, see Ref. (?). These cross-sections can still be written in terms of $`k_{}`$-dependent multi-parton correlation functions (?) and will also appear in DIS final states (?). DIS will allow us to test the universality of these correlations, that is, whether such correlations extracted from p-A collisions can be used to compute e-A final states (?, ?). #### 3.3.6 The Color Glass Condensate and the Quark Gluon Plasma The CGC provides the initial conditions for nuclear collisions at high energies. The number and energy of gluons released in a heavy ion collision of identical nuclei can be simply expressed in terms of the saturation scale as (?, ?, ?) $`{\displaystyle \frac{1}{\pi R^2}}{\displaystyle \frac{dE}{d\eta }}={\displaystyle \frac{c_E}{g^2}}Q_s^3,{\displaystyle \frac{1}{\pi R^2}}{\displaystyle \frac{dN}{d\eta }}={\displaystyle \frac{c_N}{g^2}}Q_s^2,`$ (37) where $`c_E0.25`$ and $`c_N0.3`$. Here $`\eta `$ is the space-time rapidity. These simple predictions led to correct predictions for the hadron multiplicity at central rapidities in Au-Au collisions at RHIC (?, ?) and for the centrality and rapidity dependence of hadron distributions (?). However, the failure of more detailed comparisons to the RHIC jet quenching data (?) and elliptic flow data (?) suggested that final state effects are important and significantly modify predictions based on the CGC alone. The success of hydrodynamic predictions suggests that matter may have thermalized to form a quark gluon plasma (?). Indeed, bulk features of multiplicity distributions may be described by the CGC precisely as a consequence of early thermalization–leading to entropy conservation (?). Initial-state effects will be more important in heavy ion collisions at the LHC because one is probing smaller $`x`$ in the wave function. Measurements of saturation scales for nuclei at the EIC will independently corroborate equations such as Eq. 37 and therefore the picture of heavy ion collisions outlined above. Further, a systematic study of energy loss in cold matter will help constrain extrapolations of pQCD (?) used to study jet quenching in hot matter. #### 3.3.7 Proton/Deuteron-nucleus versus electron-nucleus collisions as probes of high parton densities Both p/D-A and e-A collisions probe the small $`x`$ region at high energies. Both are important to ascertain truly universal aspects of novel physics. e-A collisions, owing to the independent ”lever” arm in $`x`$ and $`Q^2`$, as well as the simpler lepton-quark vertex, are better equipped for precision measurements. For example, in e-A collisions, information about gluon distributions can be extracted from scaling violations and from photon-gluon fusion processes. In both cases, high precision measurements are feasible. In p-A collisions, one can extract gluon distributions from scaling violations in Drell-Yan and gluon-gluon and quark-gluon fusion channels such as open charm and direct photon measurements respectively. However, for both scaling violations and fusion processes, one has more convolutions and kinematic constraints in p-A than in e-A. These limit both the precision and range of measurements. In Drell-Yan, in contrast to $`F_2`$, clear scaling violations in the data are very hard to see and data are limited to $`M^2>16`$ GeV<sup>2</sup>, above the $`J/\psi `$ and $`\psi ^{}`$ thresholds. A clear difference between p/D-A and e-A collisions is in hard diffractive final states. At HERA, these constituted approximately 10$`\%`$ of the total cross section. At eRHIC, these may constitute 30-40$`\%`$ of the cross section (?, ?). Also, factorization theorems derived for diffractive parton distributions only apply to lepton-hadron processes (?). Spectator interactions in p/D-A collisions will destroy rapidity gaps. A comparative study of p/D-A and e-A collisions thus has great potential for unravelling universal aspects of event structures in high energy QCD. ## 4 Electron-Ion Collider: Accelerator Issues With the scientific interest in a high luminosity lepton-ion collider gathering momentum during the last several years, there has been a substantial effort in parallel to develop a preliminary technical design for such a machine. A team of physicists from BNL, MIT-Bates, DESY and the Budker Institute have developed a realistic design (?) for a machine using RHIC, which would attain an e-p collision luminosity of $`0.4\times 10^{33}\mathrm{cm}^2\mathrm{s}^1`$ and could with minimal R&D start construction as soon as funding becomes available. Other more ambitious lepton-ion collider concepts which would use a high intensity electron linac to attain higher luminosity are under active consideration (?, ?). This section gives an overview of the activities currently underway related to the accelerator design. The physics program described above sets clear requirements and goals for the lepton-ion collider to be a successful and efficient tool. These goals include: a sufficiently high luminosity; a significant range of beam collision energies; and polarized beam (both lepton and nucleon) capability. On the other hand, to be realistic, the goals should be based on the present understanding of the existing RHIC machine and limitations which arise from the machine itself. Realistic machine upgrades should be considered to overcome existing limitations and to achieve advanced machine parameters, but those upgrades should be cost-effective. The intent to minimize required upgrades in the existing RHIC rings affects the choice of parameters and the set of goals. For example, the design assumed simultaneous collisions of both ion-ion and lepton-ion beams. In the main design line, collisions in two ion-ion interaction regions, at the “6” and “8 o’clock” locations, have to be allowed in parallel with electron-ion collisions. Taking these considerations into account, the following goals were defined for the accelerator design: * The machine should be able to provide beams in the following energy ranges: for the electron accelerator, 5-10 GeV polarized electrons, 10 GeV polarized positrons; for the ion accelerator, 50-250 GeV polarized protons, 100 GeV/u Gold ions. * Luminosity: in the $`10^{32}10^{33}\mathrm{cm}^2\mathrm{s}^1`$ range for e-p collisions; in the $`10^{30}10^{31}\mathrm{cm}^2\mathrm{s}^1`$ range for e-Au collisions * 70% polarization for both lepton and proton beams * Longitudinal polarization in the collision point for both lepton and proton beams An additional design goal was to include the possibility of accelerating polarized ions, especially polarized <sup>3</sup>He ions. ### 4.1 eRHIC: ring-ring design The primary eRHIC design centers on a 10 GeV lepton storage ring which intersects with one of the RHIC ion beams at one of the interaction regions (IRs), not used by any of the ion-ion collision experiments. RHIC uses superconducting dipole and quadrupole magnets to maintain ion beams circulating in two rings on a 3834 meter circumference. The ion energy range covers 10.8 to 100 GeV/u for gold ions and 25 to 250 GeV for protons. There are in total 6 intersection points where two ion rings, Blue and Yellow, cross each other. Four of these intersections points are currently in use by physics experiments. A general layout of the ring-ring eRHIC collider is shown in Figure 29 with the lepton-ion collisions occurring in the “12 o’clock” interaction region. Plans have been made for a new detector, developed and optimized for electron-ion collision studies, to be constructed in that interaction region. The electron beam in this design is produced by a polarized electron source and accelerated in a linac injector to energies of 5 to 10 GeV. To reduce the injector size and cost, the injector design includes recirculation arcs, so that the electron beam passes through the same accelerating linac sections multiple times. Two possible linac designs, superconducting and normal conducting, have been considered. The beam is accelerated by the linac to the required collision energy and injected into the storage ring. The electron storage ring is designed to be capable of electron beam storage in the energy range of 5 to 10 GeV with appropriate beam emittance values. It does not provide any additional acceleration for the beam. The electron ring should minimize depolarization effects in order to keep the electron beam polarization lifetime longer than the typical storage time of several hours. The injector system also includes the conversion system for positron production. After production the positrons are accelerated to 10 GeV energy and injected into the storage ring similarly to the electrons. Obviously the field polarities of all ring magnets should be reversed in the positron operation mode. Unlike electrons, the positrons are produced unpolarized and have to be polarized using radiative self-polarization in the ring. Therefore, the design of the ring should allow for a sufficiently small self polarization time. The current ring design provides a self polarization time of about 20 min at 10 GeV. But with polarization time increasing sharply as beam energy goes down the use of a polarized positron beam in the present design is limited to 10 GeV energy. The design of the eRHIC interaction region involves both accelerator and detector considerations. Figure 29 shows the electron accelerator located at the “12 o’clock” region. Another possible location for the electron accelerator and for electron-ion collisions might be the “4 o’clock” region. For collisions with electrons the ion beam in the RHIC Blue ring will be used, because the Blue ring can operate alone, even with the other ion ring, Yellow, being down. The interaction region design provides for fast beam separation for electron and Blue ring ion bunches as well as for strong focusing at the collision point. In this design, the other (Yellow) ion ring makes a 3m vertical excursion around the collision region, avoiding collisions both with electrons and the Blue ion beam. The eRHIC interaction region includes spin rotators, in both the electron and the Blue ion rings, to produce longitudinally polarized beams of leptons and protons at the collision point. The electron cooling system in RHIC (?, ?) is one of the essential upgrades required for eRHIC. The cooling is necessary to reach the luminosity goals for lepton collisions with Gold ions and low (below 150 GeV) energy protons. Electron cooling is considered an essential upgrade of RHIC to attain higher luminosity in ion-ion collisions. In addition, the present eRHIC design assumes a total ion beam current higher than that being used at present in RHIC operation. This is attained by operating RHIC with 360 bunches. The eRHIC collision luminosity is limited mainly by the maximum achievable beam-beam parameters and by the interaction region magnet aperture limitations. To understand this, it is most convenient to use a luminosity expression in terms of beam-beam parameters ($`\xi _e\xi _i`$) and rms angular spread in the interaction point ($`\sigma _{xi}^{},\sigma _{ye}^{}`$): $`L=f_c{\displaystyle \frac{\pi \gamma _i\gamma _e}{r_ir_e}}\xi _{xi}\xi _{ye}\sigma _{xi}^{}\sigma _{ye}^{}{\displaystyle \frac{(1+K)^2}{K}}`$ The $`f_c=28.15`$ MHz is a collision frequency, assuming 360 bunches in the ion ring and 120 bunches in the electron ring. The parameter $`K=\sigma _y/\sigma _x`$ presents the ratio of beam sizes in the interaction point. One of the basic conditions which defines the choice of beam parameters is a requirement on equal beam sizes of ion and electron beams at the interaction point: $`\sigma _{xe}=\sigma _{xi}`$ and $`\sigma _{ye}=\sigma _{ye}`$. The requirement is based on the operational experience at the HERA collider and on the reasonable intention to minimize the amount of one beam passing through the strongly nonlinear field in the outside area of the counter-rotating beam. According to the above expression, the luminosity reaches a limiting value at the maximum values of beam-beam parameters, or at the beam-beam parameter limits. For protons (and ions) the total beam-beam parameter limit was assumed to be 0.02 , following the experience and observation from other proton machines as well as initial experience from RHIC operation. With three beam-beam interaction points, two for proton-proton and one for electron-proton collisions, the beam-beam parameter per interaction point should not exceed 0.007. For the electron (or positron) beam a limiting value of the beam-beam parameter has been put at 0.08 for 10 GeV beam energy, following the results of beam-beam simulations, as well as from the experience at electron machines of similar energy range. Because the beam-beam limit decreases proportionally with the beam energy, the limiting value for 5 GeV is reduced to 0.04. The available magnet apertures in the interaction region also put a limit on the achievable luminosity. The work on the interaction region design revealed considerable difficulties to provide an acceptable design for collisions of round beams. The IR has been designed to provide low beta focusing and efficient separation of elliptical beams, with beam size ratio $`K=1/2`$. The main aperture limitation comes from the septum magnet, which leads to the limiting values of $`\sigma _{xp}^{}=93\mu `$rad. Another limitation which must be taken into account is a minimum acceptable value of the beta-function at the interaction point ($`\beta ^{}`$). With the proton rms bunch length of 20 cm, decreasing $`\beta ^{}`$ well below this number results in a luminosity degradation due to the hour-glass effect. The limiting value $`\beta ^{}=19`$ cm has been used for the design, which results in a luminosity reduction of only about 12%. A bunch length of 20 cm for Au ions would be achieved with electron cooling. Tables 1 and 2 show design luminosities and beam parameters. The positron beam intensity is assumed to be identical to the electron beam intensity, hence the luminosities for collisions involving a positron beam are equivalent to electron-ion collision luminosities. To achieve the high luminosity in the low energy tune in Table 1, the electron cooling has to be used to reduce the normalized transverse emittance of the lower energy proton beam to $`5\pi `$mm$``$mrad. Also, in that case the proton beam should have collisions only with the electron beam. Proton-proton collisions in the other two interaction points have to be avoided to allow for a higher proton beam-beam parameter. The maximum luminosity achieved in the present design is $`4.4\times 10^{32}\mathrm{cm}^2\mathrm{s}^1`$ in the high energy collision mode (10 GeV leptons on 250 GeV protons). Possible paths to luminosities as high as $`10^{33}\mathrm{cm}^2\mathrm{s}^1`$ are being explored, with studies planned to investigate the feasibility of higher electron beam intensity operation. To achieve and maintain the Au normalized transverse beam emittances shown in Table 2, electron cooling of the Au beam will be used. For the lower energy tune of electron-Gold collisions, the intensity of the Gold beam is considerably reduced because of the reduced value of the beam-beam parameter limit for the electron beam. ### 4.2 eRHIC: linac-ring design A linac-ring design for eRHIC is also under active consideration. This configuration uses a fresh electron beam bunch for each collision and so the tune shift limit on the electron beam is removed. This provides the important possibility to attain significantly higher luminosity (up to $`10^{34}\mathrm{cm}^2\mathrm{s}^1`$) than the ring-ring design. A second advantage of the linac-ring design is the ability to reverse the electron spin polarization on each bunch. A disadvantage of the linac-ring design is the inability to deliver polarized positrons. The realization of the linac beam is technically challenging and the polarized electron source requirements are well beyond present capabilities (?). Figure 30 shows a schematic layout of a possible linac-ring eRHIC design. A 450 mA polarized electron beam is accelerated in an Energy Recovery Linac (ERL). After colliding with the RHIC beam in as many as four interaction points, the electron beam is decelerated to an energy of a few MeV and dumped. The energy thus recovered is used for accelerating subsequent bunches to the energy of the experiment. ### 4.3 Other lepton-ion collider designs: ELIC A very ambitious electron-ion collider design seeking to attain luminosities up to $`10^{35}\mathrm{cm}^2\mathrm{s}^1`$ is underway at Jefferson Laboratory (?). This Electron Light Ion Collider (ELIC) design is based on use of polarized 5 to 7 GeV electrons in a superconducting ERL upgrade of the present CEBAF accelerator and a 30 to 150 GeV ion storage ring (polarized p, d, <sup>3</sup>He, Li and unpolarized nuclei up to Ar, all totally stripped). The ultra-high luminosity is envisioned to be achievable with short ion bunches and crab-crossing at 1.5 GHz bunch collision rate in up to four interaction regions. The ELIC design also includes a recirculating electron ring that would help to reduce the linac and polarized source requirements compared to the linac-ring eRHIC design of section 4.2. The ELIC proposal is at an early stage of development. A number of technical challenges must be resolved, and several R&D projects have been started. These include development of a high average current polarized electron source with a high bunch charge, electron cooling of protons/ions, energy recovery at high current and high energy, and the design of an interaction region that supports the combination of high luminosity and high detector acceptance and resolution. ## 5 Detector Ideas for the EIC The experience gained at HERA with the H1 and ZEUS detectors (?) provides useful guidance for the conceptual design of a detector that will measure a complete event ($`4\pi `$-coverage) produced in collisions of energetic electrons with protons and ions, at different beam energies and polarizations. The H1 and ZEUS detectors are general-purpose magnetic detectors with nearly hermetic calorimetric coverage. The differences between them are based on their approach to calorimetry. The H1 detector collaboration emphasized the electron identification and energy resolution, whereas the ZEUS collaboration puts more emphasis on optimizing hadronic calorimetry. The differences in their physics philosophy were reflected in their overall design: H1 had a liquid Argon calorimeter inside the large diameter magnet, whereas ZEUS chose to build a Uranium scintillator sampling calorimeter with equal response to electrons and hadrons. They put their tracking detectors inside a superconducting solenoid surrounded by calorimeters and muon chambers. H1 placed their tracking chambers inside the calorimeter surrounded by their magnet. In addition, both collaborations placed their luminosity and electron detectors downstream in the direction of the proton and electron directions, respectively. Both collaborations added low angle forward proton spectrometers and neutron detectors in the proton beam direction. Both detectors have good angular coverage (approximately, $`3^{}<\theta <175^{}`$, where the angle is measured with respect to the incoming proton beam direction) for electromagnetic (EM) calorimetry, with energy resolutions of 1-3% and for electromagnetic showers $`\sigma /E15\%/\sqrt{E(\mathrm{GeV})}+1\%`$. Hadronic energy scale uncertainties of 3% were achieved for both, with some differences in the $`\sigma /E`$ for hadronic showers, which were $`R/\sqrt{E(\mathrm{GeV})}+2\%`$, where $`R=35\%`$ and $`50\%`$ for ZEUS and H1, respectively. With the central tracking fields $`1.5`$ T covering a region similar to the calorimetric angular acceptance, momentum resolution $`\sigma /p_T<0.01p_T(\mathrm{GeV})`$ was generally achieved for almost all acceptances, except for the forward and backward directions. These directions were regarded at the beginning as being less interesting. However, the unexpected physics of low $`x`$ and low $`Q^2`$ (including diffraction in e-p scattering) came from this rather poorly instrumented region. And since the luminosity upgrade program, the low $`\beta ^{}`$ magnets installed close to the interaction point to enhance the luminosity of e-p collisions (HERA-II) have further deteriorated the acceptance of detectors in these specific geometric regions. The EIC detector design ideas are already being guided by the lessons learned from the triumphs and tribulations of the HERA experience. All advantages of the HERA detectors such as the almost 4$`\pi `$ coverage and the functionality with respect to spatial orientation will be preserved. The EIC detector will have enhanced capability in the very forward and backward directions to measure continuously the low $`x`$ and low $`Q^2`$ regions that are not comprehensively accessible at HERA. The detector design directly impacts the interaction region design and hence the accelerator parameters for the two beam elements: the effective interaction luminosity and the effective polarization of the two beams at the interaction point. Close interaction between the detector design and the IR design is hence needed in the very early stage of the project, which has already started (?). It is expected that the detector design and the IR design will evolve over the next few years. The e-p and e-A collisions at EIC will produce very asymmetric event topologies, not unlike HERA events. These asymmetries, properly exploited, allow precise measurements of energy and color flow in collisions of large and small-$`x`$ partons. They also allow observation of interactions of electrons with photons that are coherently emitted by the relativistic heavy ions. The detector for EIC must detect: the scattered electrons, the quark fragmentation products and the centrally produced hadrons. It will be the first collider detector to measure the fragmentation region of the proton or the nucleus, a domain not covered effectively at HERA. The detector design, in addition, should pose no difficulties for important measurements such as precision beam polarization (electron as well as hadron beam) and collision luminosity. The EIC detector design will allow measurements of partons from hard processes in the region around $`90^{}`$ scattering angle with respect to the beam pipes. This central region could have a jet tracker with an EM calorimeter backed by an instrumented iron yoke. Electrons from DIS are also emitted into this region and will utilize the tracking and the EM calorimetry. Electrons from photo-production and from DIS at intermediate and low momentum transfer will have to be detected by specialized backward detectors. With these guiding ideas, one could imagine that the EIC barrel might have a time projection chamber (TPC) backed by an EM Calorimeter inside a superconducting coil. One could use Spaghetti Calorimetry (SPACAL) for endcaps and GEM-type micro-vertex detector to complement the tracking capacity of the TPC in the central as well as forward/backward (endcap) regions. This type of central and end-cap detector geometry is now fairly standard. Details of the design could be finalized in the next few years using the state of the art technology and experience from more recent detectors such as BaBar at SLAC (USA) and Belle at KEK (Japan). To accommodate tracking and particle ID requirements for the different center-of-mass energy running ($`\sqrt{s}=30100`$ GeV) resulting at different beam energies, the central spectrometer magnet will have multiple field strength operation capabilities, including radial dependence of field strengths. Possible spectrometry based on dipole and toroidal fields is also being considered at this time. The forward and backward regions (hadron and electron beam directions) in e-p collisions were instrumented at HERA up to a pseudo-rapidity of $`\eta 3`$. A specialized detector added later extended this range with difficulty to $`\eta 4`$. Although acceptance enhancement in the regions beyond $`\eta =4`$ is possible with conventional ideas such as forward calorimetry and tracking using beam elements and silicon strip based Roman Pot Detectors (?), it is imperative for the EIC that this region be well instrumented. A recent detector design for eRHIC developed by the experimental group at the Max-Planck Institute, Munich, accomplishes just this (?) by allowing continuous access to physics up to $`\eta 6`$. The main difference with respect to a conventional collider detector is a dipole field, rather than a solenoid, that separates the low energy scattered electron from the beam. High precision silicon tracking stations capable of achieving $`\mathrm{\Delta }p/p2\%`$, EM calorimetry with energy resolution better than $`20\%/\sqrt{E}`$, an excellent $`e/\pi `$ separation over a large $`Q^2`$ range, all in the backward region (in the electron beam direction) are attainable. In the forward region, the dipole field allows excellent tracking and a combination of EM and hadronic calorimetry with $`20\%/\sqrt{E(\mathrm{GeV})}`$ and $`50\%/\sqrt{E(\mathrm{GeV})}`$ energy resolution, respectively. This allows access to very high $`x0.9`$ with excellent accuracy. This region of high $`x`$ is largely unexplored both in polarized and in unpolarized DIS. A significant distance away from the EIC central detector and IR, there may be Roman Pots, high rigidity spectrometers including EM calorimetry and forward electron taggers, all placed to improve the measurement of low angle scattering at high energy. Although significant effort will be made to avoid design conflicts, the conventional detector using a solenoid magnet and the one described above may not coexist in certain scenarios being considered for the accelerator designs of the EIC at BNL. The main design line, presently the ring-ring design, may be particularly difficult with only one IR. Options such as time sharing between two detectors at the same IR with the two detectors residing on parallel rails may be considered. In the case of the linac-ring scenario, several other options are available. Because the physics of low $`x`$ and low $`Q^2`$ does not require a large luminosity, nor is presently the beam polarization a crucial requirement for the physics (?), an interaction point with sufficient beam luminosity would be possible with innovative layouts of the accelerator complex. These and other details will be worked out in the next several years. Depending on the interest shown by the experimental community, accelerator designs that incorporate up to four collision points (while still allowing two hadron-hadron collision points at RHIC) will be considered and developed. ## Acknowledgments This review draws generously on the whitepaper for the Electron Ion Collider (BNL-68933-02/07-REV), and we thank all our co-authors on this publication for their efforts. We are especially grateful to Marco Stratmann and Mark Strikman for valuable advice, discussions and comments. W.V. and A.D. are grateful to RIKEN and Brookhaven National Laboratory. R.M. is supported by the Department of Energy Cooperative Agreement DE-FC02-94ER40818. W.V. and R.V. were supported by Department of Energy (contract number DE-AC02-98CH10886). R.V. ’s research was also supported in part by a research award from the A. Von Humboldt Foundation.
warning/0506/nucl-ex0506004.html
ar5iv
text
# The Thorium Molten Salt Reactor : Moving On from the MSBR ## Introduction In order to reduce $`\text{CO}_2`$ emissions in the coming decades, and, as a result, to mitigate global warming, it appears necessary to stabilize or, better, to reduce the use of fossil fuels. Resorting to a sustainable version of nuclear power may help replace classical energy production partially and thus satisfy an increasing world energy demand while conserving the climate and natural resources. The Generation IV International Forum for the development of new nuclear energy systems has established a set of goals as research directions for nuclear systems: enhanced safety and reliability, reduced waste generation, effective use of uranium or thorium ores, resistance to proliferation, improved economic competitiveness. Molten Salt Reactors (MSR) are one of the systems retained by Generation IV. MSRs are based on a liquid fuel, so that their technology is fundamentally different from the solid fuel technologies currently in use. Some of the advantages specific to MSRs (in terms of safety/reliability, for example) originate directly from this characteristic . Furthermore, this type of reactor is particularly well adapted to the thorium fuel cycle (Th-$`{}_{}{}^{233}\text{U}`$) which has the advantage of producing less minor actinides than the uranium-plutonium fuel cycle ($`{}_{}{}^{238}\text{U}`$-$`{}_{}{}^{239}\text{Pu}`$) . Moreover, while breeding or regeneration in the U-Pu cycle can be obtained only with a fast neutron spectrum, in the Th-$`{}_{}{}^{233}\text{U}`$ fuel cycle it can, in principle, be obtained with a more or less moderated neutron spectrum. In a thermal neutron spectrum, poisoning due to the Fission Products (FP) being worse than in a fast neutron spectrum, the rate at which fuel reprocessing is performed can become a major issue. Because, in an MSR, the fuel is liquid, continuous extraction of the FPs is a possibility. Although MSRs can be operated as incinerators they will be discussed in this paper only as electricity producing critical systems. In 1964, the Molten Salt Reactor Experiment (MSRE) was initiated at the Oak Ridge National Laboratory (ORNL). Generating 8 MWth of power, the reactor was operated without problems and with different fuels ($`{}_{}{}^{235}\text{U}`$ then $`{}_{}{}^{233}\text{U}`$) over several years. The expertise gained during this experiment led, in the 1970s, to the elaboration of a power reactor project, the Molten Salt Breeder Reactor (MSBR ). The studies demonstrated that fuel regeneration is possible with the thorium fuel cycle in an epithermal spectrum, provided very efficient and, as a consequence, constraining, on-line chemical reprocessing of the salt is achieved. Over the past few years, the MSBR has been reassessed in the light of new calculating methods so as to elaborate a new reactor concept that we call the Thorium Molten Salt Reactor (TMSR). The MSBR suffered from several major drawbacks and was discontinued. The goal being, at the time, to obtain as high a breeding ratio as possible, the on-line chemical reprocessing unit considered had to process the entire salt volume within 10 days and this was very complex . Because of this complexity, the project is often considered unfeasible. In addition, recent calculations have shown that the global feedback coefficient for this system is slightly positive. This contradicts the results that had been presented. The difference is probably due to the fact that, at the time, the compositions were handled in a homogeneous way while they are handled heterogeneously today. This critical issue makes the MSBR a potentially unstable system in some situations. The aim of this paper is to present solutions to these problems. In our search for better reactor configurations, we have identified several constraints that are discussed in the first part of this work. We then discuss the impact of the various reactor parameters on these constraints, i.e. chemical reprocessing, channel size, fuel volume and the proportion of heavy nuclei (HN) in the salt. A synthesis of these studies is set forth in the last section. This work is based on the coupling of a neutron transport code (MCNP ) with a materials evolution code. The former calculates the neutron flux and the reaction rates in all the cells while the latter solves the Bateman equations for the evolution of the materials composition in the cells. These calculations take into account the input parameters (power released, criticality level, chemistry,…), by adjusting the neutron flux or the materials composition of the core on a regular basis. Our calculations are based on a precise description of the geometry and consider several hundreds of nuclei with their interactions and radioactive decay; they allow fine interpretation of the results. All the data discussed in this paper result from the evolution of the reactor over 100 years. ## 1 Constraints We identify five major constraints in this study: safety, chemical reprocessing feasibility, fuel regeneration capability, materials life span, and initial inventory. Other constraints could be considered, such as waste minimization, thermal-hydraulics, or proliferation resistance but we concentrate essentially on the above five major constraints. We seek to understand the impact of the reactor’s defining parameters on these constraints. In so doing, we can single out the best reactor configurations according to the weight assigned to each of the constraints. ### 1.1 Safety In the work we present here this constraint concerns essentially the evolution of the feedback coefficients that should be negative. The more kinetic aspects of the reactor’s safety properties are not considered here. Additionally, the ways in which we change the concept do not modify the other MSR safety properties, such as fuel dumping and the fact that MSRs are free of high pressure areas. The feedback coefficients, $`\frac{dk}{dT}`$, are a measure of the variation of the multiplication factor ($`dk`$) with the temperature of the core or of a portion of the core ($`dT`$). The global feedback coefficient can be broken up into several strongly uncorrelated partial coefficients, each of which characterizes the variation of a specific parameter: the effects due to the expansion of the salt<sup>1</sup><sup>1</sup>1 The heating of the salt induces a widening of the resonances due to the Doppler effect and a change in the neutron spectrum moderation due to the salt. Both of these effects are considered together, under the term “Doppler”., and the purely thermal effects of the salt and of the graphite. This reads: $$\frac{dk}{dT})_{total}=\frac{dk}{dT})_{density}+\frac{dk}{dT})_{Doppler}+\frac{dk}{dT})_{graphite}$$ In order for the reactor to be intrinsically safe, a temperature increase must not induce an increase of the reactivity and, as a consequence, of the power released by fissions. For this reason, the $`\frac{dk}{dT})_{total}`$ coefficient must be negative. The thermal kinetics of the graphite, which is heated by gamma radiation and cooled by the salt, is much slower than that of the salt. Making allowance for this delay between the heating of the salt and the heating of the graphite, the coefficient for the salt alone, that is the sum $`\frac{dk}{dT})_{density}+\frac{dk}{dT})_{Doppler}`$ must also be negative. The degree of safety can be further increased if the density coefficient is made negative. This implies that any local loss of density, e.g. because of a bubble, decreases the reactivity of the system. The uncertainties on these values are related to statistical errors that are well identified and can be reduced, but also to systematic errors that are not quantified and are related, for example, to uncertainties in cross section evaluations. For this reason, the feedback coefficients must be sufficiently negative to ensure unambiguous stability. ### 1.2 Feasibility of the Chemical Reprocessing The term feasibility reflects the complexity associated to the chemical reprocessing. Indeed, some of the separation processes are considered too difficult to be implemented. This can have several causes: the processes considered are not well understood or mastered, the flow of materials to be processed is too large, the reprocessing implies direct coupling to the reactor core,… The objective is to devise the simplest possible system that is compatible with the other constraints. In particular, it will be important to avoid excessive deterioration of the system’s fuel regeneration capability. ### 1.3 Fuel Regeneration Capability The breeding ratio expresses the balance between the creation of $`{}_{}{}^{233}\text{U}`$ through neutron capture on $`{}_{}{}^{232}\text{Th}`$ and the destruction of $`{}_{}{}^{233}\text{U}`$ through fission or neutron capture. The breeding ratio in a critical reactor can thus be written: $$BR=\frac{r_{c,^{232}Th}r_{c,^{233}Pa}}{r_{f,^{233}U}+r_{c,^{233}U}}$$ With $`r_c`$ and $`r_f`$ respectively the capture rate and the fission rate of the different isotopes. A breeding ratio less than 1 implies that $`{}_{}{}^{233}\text{U}`$ is consumed so that fissile matter must be fed into the core on a regular basis. This inevitably increases both the volume and the frequency of transfer of these dangerous materials. Similarly, a breeding ratio larger than 1 implies that the excess $`{}_{}{}^{233}\text{U}`$ produced be placed in storage and/or transported. Because, in all cases, the initial fissile matter inventory has to be produced by other means (e.g. in pressurized water reactors or fast neutron reactors) the highest possible breeding ratio does not necessarily have to be sought. In order to satisfy the regeneration constraint, we try to achieve a breeding ratio at least equal to 1, knowing that any excess neutrons can always be put to use (improved safety, transmutation capabilities, … ). ### 1.4 Materials Life Span This concerns in particular how the graphite reacts to irradiation exposure. Beyond a certain degree of damage, it becomes the seat of swelling. Graphite’s life span is determined by the time it takes to reach a fluence limit, that we will set to $`2.10^{22}`$ n/$`\text{cm}^2`$ at a temperature of 630 °C . In our calculation, we consider only the neutrons whose energy is larger than 50 keV, i.e. those that create real damage in the graphite. The goal, with this constraint, is to obtain a life span that is not too short so as to avoid replacing the core graphite too frequently. ### 1.5 Initial Inventory The inventory, here, is the amount of $`{}_{}{}^{233}\text{U}`$ needed to start a 1 GWe power reactor. The smaller the inventory, the faster the deployment of a fleet of such reactors can be achieved . Without excluding configurations with a large inventory, its minimization will be sought. These constraints are not all equivalent; a weighting factor can be assigned to each of them. This factor depends on the technologies available and the goals that guide reactor choices. As the performance of a system depends on how the constraints are weighted and on how difficult it is to satisfy them, it is not possible to specify the “best” solution. The only possibility is to identify a number of interesting trends. This yields a better understanding of the system and can lead to the definition of a power reactor (stringent constraints) or of a demonstration unit (less stringent constraints). ## 2 Impact of the Parameters on the Constraints In this section, we examine how various reactor parameters impact the five constraints discussed above. In order to be able to compare the systems studied, we found it useful to define a standard system from which the different studies could stem. Our standard system is a 1 GWe graphite moderated reactor. Its operating temperature is 630 °C and its thermodynamic efficiency is 40 %. The graphite matrix comprises a lattice of hexagonal elements with 15 cm sides. The total diameter of the matrix is 3.20 m. Its height is also 3.20 m. The density of this nuclear grade graphite is set to 1.86. The salt runs through the middle of each of the elements, in a channel whose radius is 8.5 cm. One third of the 20 $`\text{m}^3`$ of fuel salt circulates in external circuits and, as a consequence, outside of the neutron flux. A thorium and graphite radial blanket surrounds the core so as to improve the system’s regeneration capability. The properties of the blanket are such that it stops approximately 80 % of the neutrons, thus protecting external structures from irradiation while improving regeneration. We assume that the $`{}_{}{}^{233}\text{U}`$ produced in the blanket is extracted within a 6 month period. The salt used is a binary salt, LiF - $`\text{(HN)F}_4`$, whose $`\text{(HN)F}_4`$ proportion is set at 22 % (eutectic point), corresponding to a melting temperature of 565 °C. The salt density at 630 °C is set at 4.3 with a dilatation coefficient of $`10^3`$/°C . We assume that helium bubbling in the salt circuit is able to extract the gaseous fission products and the noble metals within 30 seconds. The standard reprocessing we consider is the delayed reprocessing of the total salt volume over a 6 month period with external storage of the Pa and complete extraction of the FPs and the TRansUranians (TRU) (Figure 1). ### 2.1 Influence of the Reprocessing #### 2.1.1 How Slow Delayed Reprocessing Works As previously stated, the MSBR reprocessing is considered too complex to be feasible in the next few decades. The effectiveness of this reprocessing rested mainly on the extraction and storage of the protactinium away from the neutron flux so as to avoid, insofar as possible, the production of $`{}_{}{}^{234}\text{U}`$ by neutron capture. The half-life of $`{}_{}{}^{233}\text{Pa}`$ is 27 days and its extraction has to be markedly faster if it is to be efficient. That is why the reprocessing of the total core volume in 10 days was contemplated. The difficult part of the reprocessing is Fission Product extraction in the presence of thorium. The idea, with slow reprocessing, is to first extract the thorium, so as to avoid being handicapped by its presence in the FP extraction process. This method could not be applied in the MSBR because of the large thorium flow involved, reaching several tons per day while it is only a few hundreds of kilograms per day in the case of a six month reprocessing time. In addition, with slow reprocessing, the nuclear core can be disconnected from the processing unit, small amounts of the salt being processed individually, instead of resorting to continuous on-line reprocessing, as in the MSBR. This is a source of simplification, it allows easier control of the procedure while making the core less sensitive to possible problems in the reprocessing unit. Figure 1 gives a general view of what slow reprocessing could entail. Some of the stages shown in this general schematic, such as protactinium storage, can be eliminated while maintaining the primary assets of the reprocessing. Likewise, the neptunium extracted in the course of the first fluorination, and the other TRansUranians can be either reinjected in the core or managed separately. The advantage, in the first option, is that an “incinerating” configuration is obtained, insofar as all the TRUs are kept in the core. With the second option, the production of americium, curium, and other heavier elements is significantly reduced. The time allocated to cleaning the salt and reinjecting it can be extended considerably. Indeed, if the time needed to reprocess the core volume is equal to the time before reinjecting the salt, there is as much salt outside the core as inside it. Thus, up to 6 months can separate the extraction of the fuel salt and its reinjection in the core, after removal of the FPs. The fissile matter inventory is not increased, however, thanks to the possibility of extracting the uranium during a preliminary fluorination stage. In the case of slow reprocessing, we assume very good extraction efficiencies (they are set to 1 in the calculations) because plenty of time is available. It is too early to say that such a reprocessing scheme solves the feasibility issues; it is, however, possible to assert that the simplification of the system improves its feasibility. The impact of this reprocessing on the other constraints, in particular those of regeneration and the feedback coefficients has to be assessed. #### 2.1.2 Impact of the Reprocessing Time In Table 1, the breeding ratios obtained at equilibrium are given for various reprocessing options as applied to the reactor configuration described previously. The best breeding ratio is obtained with the MSBR reprocessing and the worst with no reprocessing other than helium bubbling in the core, and $`{}_{}{}^{233}\text{U}`$ recovery in the blanket. In the table, the MSBR reprocessing is labeled “fast (10 days)” because of the rate at which the protactinium is to be extracted. However, the extraction of the FPs is partial, making the real reprocessing rate longer (equivalent to 50 days for the FPs that capture the most). The option labeled “bubbling only” is set apart because it is dramatically different from the other configurations, making any comparison with them tricky (no equilibrium state). Varying the reprocessing time from 3 months to 2 years induces about a 0.06 loss in the breeding ratio. For these four configurations, the proportion of protactinium stored outside of the neutron flux is, respectively, 30 %, 20 %, 10 % and 5 %. However, the change in the breeding ratio is due mainly to the change in the capture rate of the FPs and, to a lesser degree, of the TRUs. On the contrary, with fast reprocessing, 80 % of the protactinium is stored outside of the neutron flux and that is the direct cause of the system’s good breeding ratio, way before the FPs and the TRUs. Thus, unless it is extracted rapidly, the Pa’s incidence on regeneration is minor. We now know the leeway afforded by the reprocessing, since a doubling of the reprocessing time induces a breeding ratio loss of about 0.02. To be precise, we should add that the degradation of the breeding ratio is three times smaller in configurations with a fast neutron spectrum, where the proportion of graphite in the core is reduced. It is important to note that the reprocessing option chosen has a moderate impact on the feedback coefficients, as shown in the table. This means that reprocessing time and safety can, in a first approximation, be considered to be independent. #### 2.1.3 Destination of the TRansUranians As previously stated, the TRUs can either be fed back into the core or they can be managed separately (incinerated in sub-critical reactors, incinerated in fast neutron reactors, or placed in storage). The choice has an impact on the regeneration capabilities, as shown in Table 2. Indeed, even if some of the TRUs fission, they impair the neutron balance because of their high capture rates. In the auto-incinerating configuration, the most capturing TRUs reach equilibrium within about 30 years and contribute to the deterioration of the neutron balance. In the same table, the influence of the TRUs on the feedback coefficients is also shown. This coefficient is slightly improved if the TRUs are kept in the core. This is because the TRUs harden the spectrum, as will be discussed further in Section 2.2. Note that, in a fast neutron spectrum configuration, the impact of TRU reinjection on both the breeding ratio and the feedback coefficients is reduced. TRU extraction, however, is advantageous in terms of waste production. When they are submitted to a neutron flux, TRUs form, progressively, significant amounts of very heavy elements such as curium. The ratio of capture to fission cross sections is not favorable to incineration in this type of reactor because of its epithermal neutron spectrum. If these elements are removed from the reactor core, larger amounts of neptunium, formed constantly by captures on $`{}_{}{}^{236}\text{U}`$, are extracted, but the production rate of the other actinides is reduced, as shown in Table 3. The goal, then, is to obtain TRUs that are more manageable in view of incorporating them in the fuel of Fast Neutron Reactors. If such an outlet for TRUs is not available, this option is of no interest. ### 2.2 Influence of the Size of the Channels The size of the channels in which the salt circulates is a fundamental parameter of the reactor. Since the size of the hexagons is kept constant in all of our studies, the size of the channels determines the moderation ratio. Changing the radius of the channels modifies the behavior of the core, placing it anywhere between a very thermalized neutron spectrum and a relatively fast spectrum. The two extreme possibilities correspond respectively to a large number of very small channels and a single big salt channel. In the latter configuration, there is no graphite in the hexagons and the core consists in a single channel. In order to allow a comparison of the results with those of the other configurations, in this case, the hexagons are treated as salt channels with an equivalent area (channel radius: 13.6 cm.). For the configurations in which the channel radius is equal to or larger than 10 cm, it is essential that the graphite of the axial reflectors be replaced with less moderating materials (e.g. zirconium carbide). Otherwise, the fissions occur massively in the vicinity of the reflectors instead of within the core. As shown in Figure 2, the radius of the channels has a strong impact on most of the constraints<sup>2</sup><sup>2</sup>2 The moderation ratio can seem to be a more universal parameter but, like the radius of the channels, it is also influenced by other parameters. An identical moderation ratio can yield very different results according to the density of the materials involved or the size of the hexagons.. #### 2.2.1 Safety Constraint The study of the feedback coefficients requires a fine analysis of the neutron spectra involved. These are shown in Figure 3 for different channel sizes up to the single channel configuration. The cross section resonances of the materials present in the core have a strong impact on the neutronic behavior of the reactor. The main resonances are visible: fission of $`{}_{}{}^{233}\text{U}`$ at about 2 eV, $`{}_{}{}^{234}\text{U}`$ capture at 5 eV, $`{}_{}{}^{232}\text{Th}`$ capture near 22 eV, and diffusion on $`{}_{}{}^{19}\text{F}`$ near 25, 50 and 100 keV. As shown in Figure 2, the total feedback coefficient becomes rather strongly negative as the spectrum hardens. This evolution is due to the conjoined variation of the three sub-coefficients, Doppler, density and graphite, as illustrated in Figure 4. The Doppler coefficient is linked to the $`{}_{}{}^{233}\text{U}`$ fission resonance and the $`{}_{}{}^{232}\text{Th}`$ capture resonance (and, to a lesser degree, to the $`{}_{}{}^{234}\text{U}`$ capture resonance). These two elements have opposite effects on the feedback coefficient: $`{}_{}{}^{233}\text{U}`$ worsens it whereas $`{}_{}{}^{232}\text{Th}`$ improves it. The thermal agitation of the salt nuclei induces a widening of these resonances so that their influence is increased. The value of the Doppler coefficient depends on how intense the flux is at these resonance values. When the spectrum hardens, the flux is more intense for high energies and less so for low energies, the thorium resonances are favored (main resonance located at 22 eV while that of $`{}_{}{}^{233}\text{U}`$ is at 2 eV). The Doppler coefficient then becomes more negative. Beyond a certain degree of spectrum hardening, the large resonances of both thorium and uranium lie in a zone where the flux has a low intensity and their importance is reduced. This explains the worsening of the Doppler coefficient for large radii. The density coefficient is related to the expansion of the salt which pushes a fraction of the fuel outside of the moderated zone. The consequence is spectrum softening because the proportion of graphite to salt is larger, thus increasing the fission rate. The effect is small for small radii where thermalization is already very efficient and where it is counterbalanced by captures in the graphite. For large radii, the thermal part of the spectrum contributes practically nothing in the neutron balance and the effects of neutron escape are felt more strongly. The density coefficient can become negative when the effects of captures in the graphite (small radius, large proportion of graphite) or of neutron escapes (large radius, fast neutron spectrum) dominate over the effects of thermalization. The graphite coefficient comes from an energy shift of the thermal part of the neutron spectrum (around 0.2 eV), due to heating of the moderator. This shift increases the fission rate because of a small low energy (0.3 eV) resonance in the fission cross section of $`{}_{}{}^{233}\text{U}`$ . Its impact on the stability decreases as the amount of graphite in the core decreases and as the influence of the thermal portion of the spectrum weakens. #### 2.2.2 Regeneration Constraint The capacity to regenerate the fuel varies a great deal with the size of the channels. This can be explained on one hand by the number of neutrons available and, on the other hand, by the increased neutron losses in configurations with small channel radii. The number of neutrons available represents the number of leftover neutrons once both the chain reaction and the regeneration are ensured. This is defined by: $$N_a=\nu 2\left(1+\alpha \right)$$ Where $`\alpha `$ is the mean capture to the mean fission cross section ratio of $`{}_{}{}^{233}\text{U}`$. These available neutrons are distributed mainly between sterile captures and supplementary captures in thorium (breeding). $`N_a`$ reaches a minimum at about r = 8.5 cm because of the variations of $`\alpha `$ with the neutron spectrum. For small radii, the strong dip in the breeding ratio is due to neutron losses in the graphite because there is so much of it. #### 2.2.3 Materials Life Span Constraint The mean cross sections decrease dramatically with the hardening of the neutron spectrum. As increasing the inventory does not compensate for this loss, the neutron flux has to be increased in order to keep the power constant. While this phenomenon is linear, this is not true of the core graphite’s life span, as shown in Figure 2. A few items have to be stressed: * The graphite in the center undergoes a flux that is more intense than in the periphery. The life span we provide is averaged over the entire core. * The maximum fluence in the graphite decreases when the temperature increases and the temperature is not uniform. Since the graphite is heated by gamma radiation and cooled by the salt, the temperature is higher between channels and lower on the channel surface. The temperature difference increases as the channels are further apart. This means that configurations with a smaller channel radius should have a smaller maximum authorized fluence than those with a larger channel radius. Generally, the flux in the graphite is directly related to the flux in the salt so that increasing the flux in the salt reduces the graphite’s life span. It is thus considerably shorter with larger channel radii than with smaller ones. As can be noticed, the graphite life span curve does not extend all the way to the single channel configuration since, in that configuration, there is no graphite inside the core (except that of the blanket). This configuration then has an asset in that it almost completely solves the issue of the graphite’s life span. #### 2.2.4 In Core Inventory Constraint For the reactor to be critical, the fissile matter inventory has to be adjusted when the neutron spectrum hardens. Indeed, the mean $`{}_{}{}^{233}\text{U}`$ fission cross section and the mean $`{}_{}{}^{232}\text{Th}`$ capture cross section decrease as the energy of the neutrons increases but the evolution is not identical for the two isotopes. Two different operating regimes can be singled out as shown in Figure 2. * For small channel radii, the cross section decreases practically in the same way for the two isotopes and the inventory required does not change much. * For larger radii, beyond 7 cm, the mean fission cross section of $`{}_{}{}^{233}\text{U}`$ decreases faster than the capture cross section of $`{}_{}{}^{232}\text{Th}`$ so that the inventory has to be increased significantly. ### 2.3 Influence of the Salt Volume The power per unit volume of salt (specific power) is a determining parameter in a reactor’s behavior. In the reference configuration, it amounts to about 250 W/$`\text{cm}^3`$ for the salt in the core. This parameter can be modified in two ways: by changing the fuel volume at fixed power or by changing the total reactor power at fixed salt volume. These two options yield similar results and only the first one is discussed in this paper. Since the flow of Heavy Nuclei is considered to be a key factor for the feasibility of the chemical reprocessing, the reprocessing time is adjusted so as to keep this flow constant from one system to the other. Thus, doubling the salt volume implies that core reprocessing takes twice as long. The incidence of the salt volume on the various constraints is rather simple, as shown in Figure 5. The size does not have a significant impact on the feedback coefficients because the neutron spectrum changes very little with the size. The slight evolution of the coefficient is due to the difference in neutron escapes, which are more likely in smaller reactors. The evolution of the breeding ratio as the salt volume increases has two main causes: the difference in neutron escapes, and the change in specific power, losses due to the Pa being a direct function of the specific power. FP capture rates do not play a significant role in this evolution. Unlike Pa, whose inventory is determined mainly by its rapid radioactive decay, their concentration at equilibrium depends on the reprocessing time. The longer reprocessing time exactly compensates the effect due to the smaller specific power. The neutron flux in the core is directly related to the specific power and the graphite’s life span varies accordingly. Similarly, the inventory in the core depends on the volume but the effect is not directly proportional, because neutron escapes are different. One last important aspect is the thermal hydraulic constraint. The thermal power is evacuated by the fuel which thus has to circulate in the exchangers. A limit has to be set on the out of core salt volume so as to allow reactor control: the delayed neutrons precursors migrate away from the neutron flux along with the salt. In our studies, the external salt volume is 1/3 of the total volume. Heat evacuation becomes more difficult as the specific power increases. Small sized or high power reactors are at a disadvantage in this respect. ### 2.4 Influence of the Salt Composition #### 2.4.1 Elimination of the Be By definition, the salt plays a central role in MSRs. Serving as the solvent for the fuel, as the moderator and as the coolant, it has to have many characteristics specific to the neutronic as well as the chemical, hydraulic, or thermal aspects. Like the MSRE, the MSBR was based on a fluoride salt, because of its good neutronic properties (capture rate and moderating capability) in a thermal spectrum. Lithium was chosen for the same reasons and beryllium because it brought the melting temperature down to 490 °C. The salt composition was 71.7% LiF \- 16% Be$`\text{F}_2`$ \- 12.3% (HN)$`\text{F}_4`$ <sup>3</sup><sup>3</sup>3 The composition that we really used in our tests with this type of salt is: 70% LiF - 17.5% Be$`\text{F}_2`$ \- 12.5% (HN)$`\text{F}_4`$.. Our first step in studying the influence of the salt composition was to eliminate beryllium from the salt bringing it to the eutectic point 78% LiF - 22% (HN)$`\text{F}_4`$. All the studies discussed up to now were done with this composition. The reasons for eliminating beryllium are based mainly on problems with its chemistry, its toxicity, and its availability. The proportion of heavy nuclei in the eutectic changes drastically and its melting point increases from 490 °C to 565 °C <sup>4</sup><sup>4</sup>4 In order to make comparisons easier and because the temperature difference is not large, the studies with the 78% LiF - 22% (HN)$`\text{F}_4`$ salt were done at the same temperature as in previous studies , i.e. 630 °C.. This temperature increase remains moderate and it seems manageable with the commonly used structure materials. Because the significantly higher proportion of Heavy Nuclei has a strong impact on the in core inventory and on the breeding ratio, we have decided to reduce the salt volume from 40 $`\text{m}^3`$ (MSBR) to 20 $`\text{m}^3`$ (reference configuration for these studies) so as to keep the same amount of Heavy Nuclei in the reactor and, as a result, similar neutronic behavior. The elimination of the beryllium impacts all five constraints. The presence of a ternary component in the salt seems to complicate the reprocessing chemistry, the risk being that this element be extracted instead of the target elements. Thus, using the LiF - (HN)$`\text{F}_4`$ salt could simplify fuel reprocessing and, as a result, bring it closer to feasibility. The impact on the other constraints is shown in Table 4. The spectrum is harder with the binary salt because of the larger proportion of HN. This translates directly into an improvement of the feedback coefficient and a larger inventory. The change in the breeding ratio is due jointly to the increased proportion of HN (positive action) and to the increased specific power (negative action). The latter also implies a neutron flux increase which, combined with the faster neutron spectrum, leads to a significant deterioration of the graphite’s life span. The elimination of beryllium has an additional and significant advantage, that is not related to the constraints that we have identified in this paper. Tritium is produced, in a system like the MSBR, by the (n,nt) reaction on $`{}_{}{}^{7}\text{Li}`$ and the (n,t) reaction on $`{}_{}{}^{6}\text{Li}`$, producing 2/3 and 1/3 respectively of the tritium . The lithium used is 99.995 % enriched with $`{}_{}{}^{7}\text{Li}`$; the $`{}_{}{}^{6}\text{Li}`$ is rapidly consumed, unless it is regenerated by an (n,$`\alpha `$) reaction on $`{}_{}{}^{9}\text{Be}`$. With the elimination of beryllium, this reaction cannot occur and, as a result, the production of tritium is reduced. #### 2.4.2 Evaluation of the LiF - (HN)$`\text{F}_4`$ ##### Temperature increase : The proportion of Heavy Nuclei in the binary salt can be adjusted. As it is reduced, the melting point increases, reaching 845 °C with pure LiF. Common structure materials cannot withstand such a temperature increase. However, new promising solutions based on carbon (carbon-carbon, carbon fiber, carbides,…) could help solve this problem . If this technology can be implemented, then the HN proportion parameter can be modified. The temperature increase due to the change of salt leads us to set the operating temperature at 1030 °C for all the configurations. We will first study the influence of this temperature hike on the standard configuration before studying the influence of the HN proportion in the salt at 1030 °C. At this temperature, the thermodynamic efficiency is assumed to increase from 40 % to 60 % and this has an incidence on the thermal power of the reactor: 1666 MWth instead of 2500 MWth are needed to produce 1000 MWe. Similarly, the salt density decreases form 4.3 to 3.89 because of the temperature related expansion effect. The impact on the constraints of this temperature increase is detailed in Table 5. The change in salt density has a direct influence on the moderation ratio, resulting in a better thermalization of the neutron spectrum. This induces, for a channel radius of 8.5 cm, a worsening of the feedback coefficient (a behavior similar to that shown in Figure 4). Similarly, this slight thermalization leads to a larger $`{}_{}{}^{233}\text{U}`$ fission cross section and, combined with the lower salt density, a smaller necessary inventory. As for the breeding ratio, it is improved because of the reduced specific power, which has a direct incidence on parasitic captures (mainly those of FPs and Pa). The lower specific power has an incidence also on the neutron flux and, thus, on the graphite’s life span. However, at such a temperature, the fluence limit that the graphite can withstand is reduced from 2.$`\text{10}^{22}`$ n/$`\text{cm}^2`$ to $`\text{10}^{22}`$ n/$`\text{cm}^2`$ . As a result, the graphite’s life span is reduced in spite of the smaller neutron flux. ##### Influence of the proportion of Heavy Nuclei: Now that the effect of the temperature change from 630 °C to 1030 °C is known, the impact of the proportion of Heavy Nuclei can be explored. It is useful to keep the total amount of Heavy Nuclei constant, as we did when we changed the salt composition. As a consequence, salts with a smaller proportion of HN will have a larger volume. This makes these new configurations potentially interesting from the point of view of thermal power extraction. The core reprocessing time is kept at 6 months since, in that case, the flow of HN to be reprocessed is the same for all the configurations. The salt of the thorium blanket is not modified. The density and expansion coefficient of the fuel salt are crucial parameters, they are given in Table 6 . Since the graphite’s life span is directly related to the specific power and since the inventory is kept constant, these constraints are not very interesting in this part of our study. We will thus concentrate our attention on the safety and the regeneration constraints. Rather than presenting the impact of the proportion of HN for a reference configuration (r = 8.5 cm, salt volume = 20 $`\text{m}^3`$) as was done previously, we will look at the impact of the channel radii for different HN proportions. This view point will allow a better understanding of the phenomena at play. The results are shown in Figures 6 and 7. During the reprocessing outlined in Figure 1, the thorium is extracted in order to allow the extraction of the FPs from the salt. This step is a key point in the reprocessing and it is made easier if the proportion of HN in the salt is small. Moreover, when this proportion is decreased, the neutrons are scattered for a longer time before they encounter a fissile or fertile element (those that dominate neutron absorptions). This leads directly to a more thermalized neutron spectrum, as shown on Figure 8. Thus, the behavior of configurations with a small proportion of HN is similar to that of configurations with a 22% (HN)$`\text{F}_4`$ proportion but with smaller channels. This additional thermalization is the main cause of the evolution of the feedback coefficients. And the visible difference for the configurations with a single salt channel seems to be due mainly to the improvement of the Doppler sub-coefficient. Likewise, the breeding ratio curves are similar from one salt to the other but two effects are observed. The first effect is related to the thermalization change and the second to a deterioration of the breeding ratio when the HN proportion is decreased. This is due to an increased capture rate in the light elements of the salt on which the neutrons scatter for a longer time. ## 3 General Discussion The various studies that have been carried out lead to a better understanding of the way an MSR works. The search for reactor configurations, be it a demonstrator or a power generator, requires that a certain number of constraints be satisfied. To do so, the different studies discussed in this paper have to be combined and the results extrapolated. Since the parameters are not mutually independent, one has to be circumspect in this approach. Let us try to explore the possible reactor configurations. For the sake of simplicity, the parameter concerning the salt composition will not be considered in a first approach. The salt, then, is 78% LiF \- 22% (HN)$`\text{F}_4`$ and the mean temperature is 630 °C. Since the safety aspect cannot be circumvented in the design process of a nuclear reactor, we consider that this constraint is necessarily satisfied. Moreover, we consider only those configurations whose total feedback coefficient, not just the salt feedback coefficient, is negative. Except in the case where the size of the reactor is reduced dramatically, leading to a significantly increased neutron flux, the total feedback coefficient is negative only for either very thermalized or fast neutron spectra. The first option implies a small fissile matter inventory and a weak neutron flux. When submitted to such a flux, the graphite undergoes little damage and its life span is reasonably long. On the other hand, captures in the moderator deteriorate the breeding ratio significantly. If a reactor system does not need to regenerate its fuel, then this very thermalized configuration may be suitable. The faster neutron spectrum option introduces a real difficulty concerning the graphite, whose life span is then on the order of one year. There are several solutions to this problem. * Decreasing the specific power of the reactor (by increasing its size and/or decreasing the total power generated) leads directly to a decreased flux intensity and, as a consequence, extends the graphite’s life span. This, however, increases in the same proportion the per GWe fissile matter inventory, without providing a very satisfactory solution. * The absence of moderating graphite in the single salt channel configuration solves this problem, the graphite in the periphery of the core being much less irradiated. However, this option leads to the fastest neutron spectrum and, as a result, the largest fissile matter inventory (5 to 6 metric tons of $`{}_{}{}^{233}\text{U}`$). The specific power can be increased, though, in order to reduce the per GWe inventory. * It may be possible to use a material whose structure is much less sensitive to irradiation than graphite. Then the configuration space for channel radii lying between 6 and 13 cm would no longer be forbidden. Finally, the high breeding ratio obtained thanks to the fast neutron spectrum provides more leeway for the reprocessing. In particular, the time for full core reprocessing can be significantly extended. The single salt channel configuration discussed earlier can even do without any reprocessing (except for the bubbling process and uranium retrieval in the blanket) and still regenerate its fuel during the first 20 years of operation. The salt could conceivably be replaced, or entirely reprocessed after this time has elapsed. If a 6 month reprocessing time is kept, it becomes possible to do without a thorium blanket while still regenerating the fuel. This would substantially simplify the reactor design. Finally, we should note that, with a fast spectrum, re-injecting the TranUranians in the core would be interesting, both in terms of regeneration (small neutron losses) and in terms of waste production (good incineration capability). If it proves possible, increasing the temperature to above 1000 °C has many positive repercussions on the constraints, in particular thanks to the increased thermodynamic efficiency. However, it induces better thermalization by the salt and this has to be taken into account. This opens the way to salt compositions containing small amounts of Heavy Nuclei. The way these reactors behave is practically the same with such compositions as with 22% (HN)$`\text{F}_4`$. In particular, the neutron spectra have to be either very thermalized or fast to ensure negative feedback coefficients. However, this corresponds to different reactor configurations because of the increased thermalization due to the salt’s light nuclei. The capture rate of these light elements deteriorates the breeding ratio and leaves less leeway than with compositions containing more HN. The fundamental importance of this parameter lies elsewhere. Indeed, it is possible either to keep the size constant, thus reducing the inventory but not the specific power, or to increase the size, thus decreasing the specific power but not the inventory. A compromise between these two extremes can be found since decreasing the specific power facilitates the evacuation of the thermal power, a constraint that must not be neglected. ## Conclusion While our studies were, at first, close to the MSBR configuration, they prompted us to diversify our investigations. We analyzed the impact on the behavior of the core of such parameters as the reprocessing, the moderation ratio, the core size, and the proportion of heavy nuclei in the salt. Our results confirm that there is a problem with the feedback coefficients in the MSBR. In a thermal spectrum, it would be possible to reach an acceptable concept only after in depth investigations taking into account the effect of the salt (negative feedback coefficient) and of the graphite (which makes the global feedback coefficient positive) separately. For very thermalized spectra, the global coefficients are negative thanks to the large neutron losses in the moderator but this leads also to a very poor breeding ratio. Epithermal or fast neutron spectra thus seem more favorable since they combine good feedback coefficients with satisfactory breeding ratio. However they lead to severe problems with the graphite’s ability to withstand the irradiation. As a result, the solution that removes the moderating block seems especially attractive. Our studies have uncovered a wider range of possibilities than anticipated. Thus, many options remain to be explored. In particular, the evaluation of new materials, be it to obtain a moderator that has better irradiation resistance properties or to allow high temperature operation is crucial for an even more interesting development of the concept. How to extract the thermal power from the core is another issue of major interest since it impacts the behavior of the core through the specific power aspect. In order to ease heat recovery, the salt composition can also be modified so as to dissolve the fissile matter in a larger salt volume. In general, it is possible to change the type of salt, the MSR concept being adaptable to such a change. As many parameters remain to be studied, other acceptable solutions could be found. In particular, parameters such as the type of salt, the moderating material, the size of the lattice hexagons, the definition of several different areas in the core, could be studied more specifically. In view of the results already obtained, it is clear that many configurations remain to be explored, requiring research on the salt and the materials as well as on the neutronics and the geometry. ## Acknowledgements The paper took benefit from the works made with M.Allibert and discussions during the european contract MOST (MOlten SalT : FIKW-CT-2001-20183). We would like to thank Elisabeth Huffer for her translation (from French to English) of this paper.
warning/0506/hep-ph0506292.html
ar5iv
text
# 2005 International Linear Collider Workshop - Stanford, U.S.A. Effects of Finite Top Lifetime at the 𝑡⁢𝑡̄ Threshold ## I Introduction Because in the Standard Model the top quark width $`\mathrm{\Gamma }_t1.5`$ GeV is much larger than the typical hadronization energy $`\mathrm{\Lambda }_{\mathrm{QCD}}`$, the total cross section $`\sigma (e^+e^{}t\overline{t})`$ in the threshold region $`\sqrt{s}2m_t`$ can be computed with perturbative methods to very high precision. The rapid rise of the cross section in the threshold region will allow for a measurement of the top quark mass (in a threshold mass scheme synopsis ) with experimental and theoretical uncertainties at the level of only $`100`$ MeV. Other parameters such as $`\alpha _s`$, $`\mathrm{\Gamma }_t`$ or (if the Higgs boson is light) the top Yukawa coupling $`g_{\mathrm{tth}}`$ can also be determined if the normalization and the exact form of the line-shape can be computed with small theoretical errors at the percent level TTbarsim ; synopsis . This is required because the observable cross section is a convolution of the theory prediction with the partly machine-dependent $`e^+e^{}`$ luminosity spectrum TTbarsim . The theoretical instrument to make first principles predictions for top threshold observables is based on NRQCD, an effective theory (EFT) of QCD for heavy quark pairs with small relative velocity $`v\alpha _s1`$. Within “velocity” NRQCD (vNRQCD) LMR ; HoangStewartultra , which we will use here, it is possible to sum Coulomb singularities and (using renormalization group methods) logarithms of $`v`$ to all orders of QCD perturbation theory while a systematic and coherent $`v`$ power counting is maintained. The summation of logarithms avoids large normalization uncertainties hmst ; hmst1 that were obtained in earlier fixed-order predictions synopsis . At NNLL order (i.e. including corrections of order $`v^2`$ and the summation of terms $`\alpha _s^2(\alpha _s\mathrm{ln}v)^n`$) all QCD ingredients for the threshold cross section are presently known except for still missing subleading mixing effects in the running of the heavy quark pair production current. The current normalization QCD uncertainties of the total cross section are estimated to be around $`6\%`$ HoangEpiphany . In this talk we are interested in electroweak effects. At leading order, the three basic electroweak effects are the $`t\overline{t}`$ production process itself, the finite top lifetime and the luminosity spectrum mentioned above. The latter is accounted for in the experimental simulations TTbarsim , the two former effects are described by NRQCD. Here we want to discuss finite top lifetime effects at the subleading level including QCD interference effects. Previous partial analyses have indicated that the corrections can reach the level of a few percent GuthKuehn ; HoangTeubnerdist . In particular, we investigate the role of absorptive parts related to the top quark decay in electroweak loop corrections to the NRQCD matching conditions that contribute to the NNLL total cross section. Interestingly, these corrections lead to UV phase space divergences that would not exist for stable top quarks and that lead to a NLL renormalization of $`(e^+e^{})(e^+e^{})`$ operators that also contributes to the total cross section. For details of our analysis we refer the interested reader to Ref. HoangReisser . ## II Power Counting and Matching Conditions Let us first recall the power counting to classify the order at which electroweak effects can contribute by considering the matching conditions for the vNRQCD bilinear quark field operators. Electroweak corrections are obtained by matching 2-point functions in the effective theory to those in QCD and the electroweak theory. The NNLL result including also all QCD contributions has the form HoangTeubnerdist $`_{\mathrm{bilinear}}(x)`$ $`=`$ $`{\displaystyle \underset{𝒑}{}}\psi _𝒑^{}(x)\left\{iD^0{\displaystyle \frac{(𝒑i𝐃)^2}{2m_t}}+{\displaystyle \frac{𝒑^4}{8m_t^3}}+{\displaystyle \frac{i}{2}}\mathrm{\Gamma }_t\left(1{\displaystyle \frac{𝒑^2}{2m_t^2}}\right)\delta m_t\right\}\psi _𝒑(x)+(\psi _𝒑\chi _𝒑),`$ (1) where the fields $`\psi _𝒑`$ and $`\chi _𝒑`$ destroy top and antitop quarks with momentum $`𝒑`$ and positive energies, $`D^\mu =(D^0,𝐃)=^\mu +igA^\mu `$ is the ultrasoft gauge covariant derivative and $`\mathrm{\Gamma }_t`$ is the top quark width defined at the (electroweak) top quark pole. The $`v`$-counting is $`D^0m_tv^2\mathrm{\Gamma }_tm_tg^2`$ since the top width is of order of the typical top kinetic energy, which defines the ultrasoft scale $`m_tv^2`$. This leads to the counting $`v\alpha _sgg^{}`$ for the SU(2) and U(1) gauge couplings $`g`$ and $`g^{}`$, where we can treat the weak mixing being of order one. The term $`\delta m_t`$ is a residual mass term of order $`v^2`$ that arises within threshold mass schemes synopsis . Its electroweak contributions are straightforward to compute and will not be discussed here further. The term $`i\mathrm{\Gamma }_t(1𝐩^2/2m_t^2))`$ describes the finite lifetime at LL order and the NNLL time dilatation effects. It produces the known replacement rule $`EE+i\mathrm{\Gamma }_t`$ Fadin1 to account for finite lifetime effects at LL order. Although it renders the effective Lagrangian formally non-hermitian (since hermitian conjugation does not change its sign), the EFT inherits unitarity from the underlying theory, so we can later use the optical theorem to compute the total rates. It is crucial to understand that the effects of the top decay represent hard physics that can be integrated out (i.e. treated by the matching condition $`i\mathrm{\Gamma }_t`$) and that, therefore, one can determine the corresponding EFT matching conditions for on-shell top quark amplitudes. This is analogous to the treatment of photons in an absorptive medium in the optical theory. So to account for all NNLL order electroweak effects to Eq. (1) one still has to include the one-loop electroweak and the $`𝒪(\alpha _s)`$ and $`𝒪(\alpha _s^2)`$ QCD corrections to the on-shell top width $`\mathrm{\Gamma }_t`$ which will also not be discussed here any further. Concerning the $`t\overline{t}`$ interaction at LL order we only need to consider the Coulomb potential, $`_{\mathrm{pot}}=_{𝒑,𝒑^{}}\frac{4\pi C_F\alpha _s(m_t\nu )}{(𝐩𝐩^{})^2}\psi _𝒑^{}^{}\psi _𝒑\chi _𝒑^{}^{}\chi _𝒑`$, where $`\nu `$ is the vNRQCD renormalization scaling parameter. Since effects from the top decay are hard ($`m_t`$), we can neglect the momentum exchange $`𝐩𝐩^{}`$ in the electroweak loops during the matching computation. From this it is easy to see that all the $`g^2`$ (vertex and wave function) corrections to the Coulomb potential cancel due to SU(3) gauge invariance. Since all electroweak corrections to the $`1/m_t`$ suppressed potentials are beyond NNLL order simply by counting powers of $`v`$ and $`g`$, we also don’t have to consider any electroweak corrections to these potentials. The same argument applies to electroweak corrections to the interactions of top quarks with the soft gluons ($`km_tv`$) or potential four-quark operators caused by electroweak box diagrams; the former can only contribute to $`𝒪(\alpha _s)`$ corrections to the potentials, and the latter are already $`1/m_t^2`$ suppressed without even accounting for the powers of $`g`$. Moreover, the Coulomb interaction does not generate any UV divergences, so electroweak NNLL contributions due to mixing do also not exist. The dominant operators used to describe $`t\overline{t}`$ spin-triplet production have the form $`𝒪_{V,𝒑}=\left[\overline{e}\gamma _je\right]𝒪_{𝒑,1}^j,𝒪_{A,𝒑}=\left[\overline{e}\gamma _j\gamma _5e\right]𝒪_{𝒑,1}^j,`$ (2) where $`𝒪_{𝒑,1}^j=\left[\psi _𝒑^{}\sigma _j(i\sigma _2)\chi _𝒑^{}\right]`$. They give the contribution $`\mathrm{\Delta }=_𝒑\left(C_V𝒪_{V,𝒑}+C_A𝒪_{A,𝒑}\right)+\text{h.c.}`$ to the effective theory Lagrangian where the hermitian conjugation (which gives the corresponding annihilation operators) is referring to the operators, but does not affect their Wilson coefficients. Since we neglect QED radiative corrections, the $`e^\pm `$ fields act like classic fields, but we need them due to electroweak gauge invariance. The leading order matching conditions to $`C_{V/A}`$ are obtained from the full theory Born diagrams with photon and Z exchange and are of order $`g^2`$. We will see in Sec. III that $`C_{V/A}`$ will receive imaginary matching conditions from one-loop electroweak corrections that contribute at NNLL order and play a role similar to the imaginary terms in Eq. (1). It has been shown in Refs. Fadin2 that up to NLL order for $`\sigma _{\mathrm{tot}}`$ there are no QCD interference effects from (ultrasoft) gluon radiation off the top/antitop quark or its decay products. (For this consideration only ultrasoft gluons are relevant because they cannot kick a top quark off-shell.) The proof was conducted by explicit computation of diagrams. Using the analysis of matching conditions we can show that this statement even holds at NNLL order. For the time-like ultrasoft $`A^0`$ gluons, QCD gauge invariance ensures that the dominant electroweak matching corrections to the $`A^0`$ interaction vertex shown in Eq. (1) vanish because we can again approximate the ultrasoft gluon momentum in the electroweak loop as zero. Moreover, the exchange of time-like ultrasoft $`A^0`$ gluons does itself not contribute at LL because they can be removed from the LL particle-antiparticle sector in Eq. (1) by a redefinition of the top and antitop fields related to static Wilson lines Korchemsky1 . An anomalous interaction in analogy to the $`g2`$ is not covered by this argument but suppressed by a factor $`1/m_t^2`$. Accounting also for powers of $`g`$ one finds that interference from $`A^0`$ gluons does not contribute at NNLL order. For the space-like ultrasoft $`𝐀`$ gluons the conclusion is the same because they couple to the quarks with the $`𝐩.𝐀/m_t`$ coupling. The $`g^2`$-suppression from an additional electroweak loop correction to the interaction vertex then leads to a contribution beyond the NNLL order. ## III Absorptive Matching Conditions To determine the absorptive parts related to the top decay of the matching conditions of the operators $`𝒪_{V,𝒑}`$ and $`𝒪_{A,𝒑}`$ that contribute to the total cross section at NNLL order, we have to consider the $`bW^+`$ and $`\overline{b}W^{}`$ cuts of the full theory diagrams shown in Fig. 1a. To obtain the contributions at this order the external (on-shell) top quarks can be taken to be at rest. The full theory amplitude has the form $`𝒜`$ $`=`$ $`i\left[\overline{v}_{e^+}(k^{})\gamma ^\mu (iC_V^{\mathrm{bW},\mathrm{abs}}+iC_A^{\mathrm{bW},\mathrm{abs}}\gamma _5)u_e^{}(k)\right]\left[\overline{u}_t(p)\gamma _\mu v_{\overline{t}}(p)\right],`$ (3) where $`k+k^{}=2p=(2m_t,0)`$. The amplitude of the charge conjugated process describing top pair annihilation reads $`\overline{𝒜}`$ $`=`$ $`i\left[\overline{u}_e^{}(k)\gamma ^\mu (iC_V^{\mathrm{bW},\mathrm{abs}}+iC_A^{\mathrm{bW},\mathrm{abs}}\gamma _5)v_{e^+}(k^{})\right]\left[\overline{v}_{\overline{t}}(p)\gamma _\mu u_t(p)\right].`$ (4) We used the cutting equations to obtain expressions for $`iC_{V/A}^{\mathrm{bW},\mathrm{abs}}`$ and checked electroweak gauge invariance by carrying out the computation in unitary and Feynman gauge. Analytic formulae are provided in Ref. HoangReisser . The results are consistent with results obtained earlier in Ref. GuthKuehn . It is a consequence of the unitarity of the underlying theory that the sign of the imaginary part of the amplitude does not change in the charge conjugated amplitude. It is straightforward to match the amplitudes for the operators $`𝒪_{V/A,𝒑}`$ and $`𝒪_{V/A,𝒑}^{}`$ to the full theory results in Eqs. (3) and (4). The resulting matching conditions for the operators $`𝒪_{V/A,𝒑}`$ and $`𝒪_{V/A,𝒑}^{}`$ read $`C_{V/A}(\nu =1)`$ $`=`$ $`C_{V/A}^{\mathrm{born}}+iC_{V/A}^{\mathrm{bW},\mathrm{abs}},`$ (5) where we have also indicated the Born level contributions. In a full treatment of electroweak effects the coefficients $`C_{V/A}`$ also include the real parts of the full set of electroweak one-loop diagrams indicated in Fig. 1a. A comprehensive examination of these contributions will be provided in subsequent analyses. ## IV Time-Ordered Product and Renormalization Through the optical theorem the NNLL order corrections to the total cross section that come from the absorptive one-loop electroweak matching conditions for the operators $`𝒪_{V/A,𝒑}`$ and from the time dilatation corrections can be computed from the imaginary part of the $`(e^+e^{})(e^+e^{})`$ forward scattering amplitude, $`\sigma _{\mathrm{tot}}`$ $``$ $`{\displaystyle \frac{1}{s}}\text{Im}\left[\left(C_V^2(\nu )+C_A^2(\nu )\right)L^{lk}𝒜_1^{lk}\right],`$ (6) where $`L^{lk}=\frac{1}{2}(k+k^{})^2(\delta ^{lk}\widehat{e}^l\widehat{e}^k)`$ is the spin-averaged lepton tensor and $`𝒜_1^{lk}=i{\displaystyle \underset{𝒑,𝒑^{\mathbf{}}}{}}{\displaystyle d^4xe^{i\widehat{q}x}\mathrm{\hspace{0.17em}0}\left|T𝒪_{𝒑,1}^{l}{}_{}{}^{}(0)𝒪_{𝒑^{\mathbf{}},1}^k(x)\right|\mathrm{\hspace{0.17em}0}}=2N_c\delta ^{lk}G^0(a,v,m_t,\nu )`$ (7) is the time-ordered product of the $`t\overline{t}`$ production and annihilation operators $`𝒪_{𝒑,1}^j`$ and $`𝒪_{𝒑,1}^{j}{}_{}{}^{}`$ hmst1 . Here we used $`k+k^{}=(\sqrt{s},0)`$, $`\widehat{𝐞}=𝐤/|𝐤|`$ and $`\widehat{q}(\sqrt{s}2m_t,0)`$, $`\sqrt{s}`$ being the c. m. energy. The result reads $$\mathrm{\Delta }\sigma _{\mathrm{tot}}^{\mathrm{\Gamma },1}=2N_c\text{Im}\left\{2i\left[C_V^{\mathrm{born}}C_V^{\mathrm{bW},\mathrm{abs}}+C_A^{\mathrm{born}}C_A^{\mathrm{bW},\mathrm{abs}}\right]G^0(a,v,m_t,\nu )+\left[(C_V^{\mathrm{born}})^2+(C_A^{\mathrm{born}})^2\right]\delta G_\mathrm{\Gamma }^0(a,v,m_t,\nu )\right\},$$ (8) where $`a𝒱_c^{(s)}(\nu )/4\pi =C_F\alpha _s(m_t\nu )`$. The term $`G^0`$ is the zero-distance S-wave Green function of the non-relativistic Schrödinger equation which is obtained from the LL terms in Eqs. (1) and the Coulomb potential. In dimensional regularization it has the form hmst1 $`G^0(a,v,m_t,\nu )={\displaystyle \frac{m_t^2}{4\pi }}\left\{iva\left[\mathrm{ln}\left({\displaystyle \frac{iv}{\nu }}\right){\displaystyle \frac{1}{2}}+\mathrm{ln}2+\gamma _E+\psi \left(1{\displaystyle \frac{ia}{2v}}\right)\right]\right\}+{\displaystyle \frac{m_t^2a}{4\pi }}{\displaystyle \frac{1}{4ϵ}},`$ (9) where $`v=\left((\sqrt{s}2(m_t+\delta m_t)+i\mathrm{\Gamma }_t)/m_t\right)^{1/2}`$. The term $`\delta G_\mathrm{\Gamma }^0`$ represents the corrections originating from the time dilatation correction in Eq. (1) and reads $`\delta G_\mathrm{\Gamma }^0(a,v,m_t,\nu )=i\frac{\mathrm{\Gamma }_t}{2m_t}[1+\frac{v}{2}\frac{}{v}+a\frac{}{a}]G^0(a,v,m_t,\nu )`$. Note that the Wilson coefficients $`C_{V/A}`$ do not have a LL anomalous dimension, so only the matching conditions at $`\nu =1`$ appear in Eq. (8). One can check that the terms proportional to $`C_{V/A}^{\mathrm{bW},\mathrm{abs}}`$ in Eq. (8) are in agreement with the full theory matrix elements from the interference between the double-resonant amplitudes for the process $`e^+e^{}t\overline{t}bW^+\overline{b}W^{}`$ (first diagram in Fig. 1b) and the single-resonant amplitudes describing the processes $`e^+e^{}t+\overline{b}W^{}bW^+\overline{b}W^{}`$ and $`e^+e^{}bW^+\overline{t}bW^+\overline{b}W^{}`$ in the $`t\overline{t}`$ threshold limit for $`m_t\mathrm{}`$ (subsequent diagrams in Fig. 1b). To find literal agreement between full and effective theory matrix elements one has to replace the $`iϵ`$ terms in the resonant full theory top propagators by the Breit-Wigner term $`im_t\mathrm{\Gamma }_t/2`$. As discussed above, there are no further QCD corrections in the non-relativistic limit due to the cancellation of the QCD interference effects caused by gluons with ultrasoft momenta. The corrections given in Eq. (8) exhibit UV $`1/ϵ`$-divergences that have interesting features. Physically they arise from a logarithmic high energy behavior of the top-antitop effective theory phase space integration for matrix elements containing a single insertion of the Coulomb potential. Technically they enter the imaginary part of the forward scattering amplitude because the imaginary parts of the matching conditions of Eqs. (5) lead to a dependence on the real part of $`G^0`$ (see Eq. (9)). In the full theory this logarithmic behavior is regularized by the top quark mass. It is important that the divergences only exist because the top quark is not treated as a stable particle. In particular, the UV divergences from the time dilatation corrections arise from the Breit-Wigner-type high energy behavior of the effective theory top propagator derived from Eq. (1), which differs from the one for a stable particle. Likewise, the interference effects described by the absorptive electroweak matching conditions for the operators $`𝒪_{V,𝒑}`$ and $`𝒪_{A,𝒑}`$ would not have to be accounted for if the top quarks were stable particles. UV divergences of the same kind for the NNLL total cross section have been observed and noted before hmst ; hmst1 ; HoangTeubnerdist ; HoangTeubner , but no resolution of the issue was provided. From the point of view of having an EFT with non-hermitian contributions it is obvious that these UV divergences must be handled in the canonical way using renormalization. Since the divergences are directly related to operators with non-hermitian Wilson coefficients, the renormalization procedure will naturally involve operators with non-hermitian Wilson coefficients. The operators that are renormalized by the UV divergences displayed in Eq. (8) are the two $`(e^+e^{})(e^+e^{})`$ operators $`\stackrel{~}{𝒪}_V=\left[\overline{e}\gamma ^\mu e\right]\left[\overline{e}\gamma _\mu e\right],\stackrel{~}{𝒪}_A=\left[\overline{e}\gamma ^\mu \gamma _5e\right]\left[\overline{e}\gamma _\mu \gamma _5e\right],`$ (10) which give the additional contribution $`\stackrel{~}{\mathrm{\Delta }}=\stackrel{~}{C}_V\stackrel{~}{𝒪}_V+\stackrel{~}{C}_A\stackrel{~}{𝒪}_A`$ to the effective theory Lagrangian, $`\stackrel{~}{C}_{V/A}`$ being the Wilson coefficients. Because in this work we neglect QED effects, the electron and positron act as classic fields and therefore $`\stackrel{~}{C}_V`$ and $`\stackrel{~}{C}_A`$ run only through mixing due to UV divergences such as in Eq. (8). Since only the imaginary parts of the coefficients $`\stackrel{~}{C}_{V/A}`$ are relevant for the discussion, we neglect the real contributions in the following. Using the standard $`\overline{\mathrm{MS}}`$ subtraction procedure to determine the (non-hermitian) counterterms of the renormalized $`\stackrel{~}{𝒪}_{V/A}`$ operators and standard methods to compute and to solve the anomalous dimensions, one obtains the following form of the Wilson coefficients $`\stackrel{~}{C}_{V/A}`$ for scales below $`m_t`$ ($`v<1`$), $`\stackrel{~}{C}_{V/A}(\nu )`$ $`=`$ $`\stackrel{~}{C}_{V/A}(1)+i{\displaystyle \frac{2N_cm_t^2C_F}{3\beta _0}}\{[((C_{V/A}^{\mathrm{born}})^2+(C_{V/A}^{\mathrm{ax}})^2){\displaystyle \frac{\mathrm{\Gamma }_t}{m_t}}+3C_{V/A}^{\mathrm{born}}C_{V/A}^{\mathrm{bW},\mathrm{abs}}]\mathrm{ln}(z)`$ (11) $`{\displaystyle \frac{4C_F}{\beta _0}}{\displaystyle \frac{\mathrm{\Gamma }_t}{m_t}}(C_{V/A}^{\mathrm{born}})^2\mathrm{ln}^2(z)+{\displaystyle \frac{4(C_A+2C_F)}{\beta _0}}{\displaystyle \frac{\mathrm{\Gamma }_t}{m_t}}(C_{V/A}^{\mathrm{born}})^2\rho (z)\},`$ where $`z\alpha _s(m_t\nu )/\alpha _s(m_t)`$, $`\rho (z)=\frac{\pi ^2}{12}\frac{1}{2}\mathrm{ln}^22+\mathrm{ln}2\mathrm{ln}(z)\mathrm{Li}_2\left(\frac{z}{2}\right)`$ and the $`\stackrel{~}{C}_{V/A}(1)`$ are the hard matching conditions. The operators $`\stackrel{~}{𝒪}_{V/A}`$ lead to an additional contribution to the total cross section of the form $`\mathrm{\Delta }\sigma _{\mathrm{tot}}^{\mathrm{\Gamma },2}`$ $`=`$ $`\text{Im}\left[\stackrel{~}{C}_V+\stackrel{~}{C}_A\right].`$ (12) For details of the computations see Ref. HoangReisser . Parametrically $`\mathrm{\Delta }\sigma _{\mathrm{tot}}^{\mathrm{\Gamma },2}`$ is of order $`g^6`$ and thus constitutes a NLL contribution as one can also see from the fact that the corresponding UV divergences were generated in NNLL order effective theory matrix elements. The correction $`\mathrm{\Delta }\sigma _{\mathrm{tot}}^{\mathrm{\Gamma },2}`$ is energy-independent, but it is scale-dependent and compensates the logarithmic scale-dependence in the NNLL contribution $`\mathrm{\Delta }\sigma _{\mathrm{tot}}^{\mathrm{\Gamma },1}`$. The matching conditions $`\stackrel{~}{C}_{V/A}(\nu =1)`$ are presently unknown and will be analysed in subsequent work. For now we set them to zero in the numerical analysis presented below. We note that it was shown in HoangTeubnerdist that the difference between the full theory phase space (which is cut off by the large, but finite $`m_t`$) and the effective theory phase space (which is infinite in the computation of the forward scattering amplitude) contributes to $`\stackrel{~}{C}_{V/A}(\nu =1)`$ and also represents a NLL effect. ## V Numerical Analysis In Fig. 2 we have plotted the sum of $`\mathrm{\Delta }\sigma _{\mathrm{tot}}^{\mathrm{\Gamma },1}`$ and $`\mathrm{\Delta }\sigma _{\mathrm{tot}}^{\mathrm{\Gamma },2}`$ in picobarn in the 1S mass scheme HoangTeubnerdist for $`M_{1\mathrm{S}}=175`$ GeV, the fine structure constant $`\alpha =1/125.7`$, $`s_w^2=0.23120`$, $`V_{tb}=1`$ and $`M_W=80.425`$ GeV with the renormalization scaling parameter $`\nu =0.1`$ (solid curves), $`0.2`$ (dashed curves) and $`0.3`$ (dotted curves). For the QCD coupling we used $`\alpha _s(M_Z)=0.118`$ as an input and employed 4-loop renormalization group running. Note that in the 1S scheme $`\delta m_t=M_{1\mathrm{S}}(𝒱_c^{(s)}(\nu )/4\pi )^2/8`$. For the top quark width we adopted the value $`\mathrm{\Gamma }_t=1.43`$ GeV. Note that in a complete analysis of electroweak effects the top quark width depends on the input parameters given above and is not an independent parameter. For the purpose of the numerical analysis in this work, however, our treatment is sufficient. We find that the sum of the corrections is negative and shows a moderate $`\nu `$ dependence. We find that the corrections are around $`10\%`$ for energies below the peak, between $`2\%`$ and $`4\%`$ close to the peak and about $`2\%`$ above the peak. Interestingly, they partly compensate the sizeable positive QCD corrections found in Hoang3loop ; HoangEpiphany . The peculiar energy dependence of the corrections, caused by the dependence on the real part of the Green function $`G^0`$, also leads to a slight displacement of the peak position. Relative to the peak position of the LL cross section one obtains a shift of $`(30,35,47)`$ MeV for $`\nu =(0.1,0.2,0.3)`$. This shift is comparable to the expected experimental uncertainties of the top mass measurements from the threshold scan TTbarsim .
warning/0506/cond-mat0506285.html
ar5iv
text
# Flux Reversal in a Two-state Symmetric Optical Thermal Ratchet ## Abstract A Brownian particle’s random motions can be rectified by a periodic potential energy landscape that alternates between two states, even if both states are spatially symmetric. If the two states differ only by a discrete translation, the direction of the ratchet-driven current can be reversed by changing their relative durations. We experimentally demonstrate flux reversal in a symmetric two-state ratchet by tracking the motions of colloidal spheres moving through large arrays of discrete potential energy wells created with dynamic holographic optical tweezers. The model’s simplicity and high degree of symmetry suggest possible applications in molecular-scale motors. Until fairly recently, random thermal fluctuations were considered impediments to inducing motion in systems such as motors. Fluctuations can be harnessed, however, through mechanisms such as stochastic resonance Hanggi (2002) and thermal ratchets Reimann (2002), as efficient transducers of input energy into mechanical motion. Unlike conventional machines, which battle noise, molecular-scale devices that exploit these processes actually requite thermal fluctuations to operate. This article focuses on thermal ratchets in which the random motions of Brownian particles are rectified by a time-varying potential energy landscape. Even when the landscape has no overall slope and thus exerts no average force, directed motion still can result from the accumulation of coordinated impulses. Most thermal ratchet models break spatiotemporal symmetry by periodically translating, tilting or otherwise modulating a spatially asymmetric landscape Reimann (2002). Inducing a flux is almost inevitable in such systems unless they satisfy conditions of spatiotemporal symmetry or supersymmetry Reimann (2001). Even a spatially symmetric landscape can induce a flux with appropriate driving Chen (1997); Kanada and Sasaki (1999); Jones et al. (2004); Lee et al. (2005). Unlike deterministic motors, however, the direction of motion in these systems can depend sensitively on implementation details. We recently demonstrated a spatially symmetric three-state thermal ratchet for micrometer-scale colloidal particles implemented with arrays of holographic optical tweezers, each of which constitutes a discrete potential energy well Lee et al. (2005). Repeatedly displacing the array first by one third of a lattice constant and then by two thirds breaks spatiotemporal symmetry in a manner that induces a flux. Somewhat surprisingly, the *direction* of motion depends sensitively on the duration of the states relative to the time required for a particle to diffuse the inter-trap separation Lee et al. (2005). The induced flux therefore can be canceled or even reversed by varying the rate of cycling, rather than the direction. This approach builds upon the pioneering demonstration of unidirectional flux induced by a spatially asymmetric time-averaged optical ratchet Faucheux et al. (1995); Harada and Yoshikawa (2004), and of reversible transitions driven by stochastic resonance in a dual-trap rocking ratchet McCann et al. (1999); Dykman and Golding (2000). Here, we demonstrate flux induction and flux reversal in a symmetric *two-state* thermal ratchet implemented with dynamic holographic optical trap arrays Dufresne and Grier (1998); Curtis et al. (2002). The transport mechanism for this two-state ratchet is more subtle than our previous three-state model in that the direction of motion is not easily intuited from the protocol. Its capacity for flux reversal in the absence of external loading, by contrast, can be inferred immediately by considerations of spatiotemporal symmetry. This also differs from the three-state ratchet Lee et al. (2005) and the rocking double-tweezer McCann et al. (1999); Dykman and Golding (2000) in which flux reversal results from a finely tuned balance of parameters. Figure 1 schematically depicts how the two-state ratchet operates. Each state consists of a pattern of discrete optical traps, modeled here as Gaussian wells of width $`\sigma `$ and depth $`V_0`$, uniformly separated by a distance $`L\sigma `$. The first array of traps is extinguished after time $`T_1`$ and replaced immediately with a second array, which is displaced from the first by $`L/3`$. The second pattern is extinguished after time $`T_2`$ and replaced again by the first, thereby completing one cycle. If the potential wells in the second state overlap those in the first, then trapped particles are handed back and forth between neighboring traps as the states cycle, and no motion results. This also is qualitatively different from the three-state ratchet, which deterministically transfers particles forward under comparable conditions, in a process known as optical peristalsis Koss and Grier (2003); Lee et al. (2005). The only way the symmetric two-state ratchet can induce motion is if trapped particles are released when the states change and then diffuse freely. The motion of a Brownian particle in this system can be described with the one-dimensional Langevin equation $$\eta \dot{x}(t)=V^{}(x(t)f(t))+\xi (t),$$ (1) where $`\eta `$ is the fluid’s dynamic viscosity, $`V(x)`$ is the potential energy landscape, $`V^{}(x)=V(x)/x`$ is its derivative, and $`\xi (t)`$ is a delta-correlated stochastic force representing thermal noise. The potential energy landscape in our system is spatially periodic with period $`L`$, $$V(x+L)=V(x).$$ (2) The time-varying displacement of the potential energy in our two-state ratchet is described by a periodic function $`f(t)`$ with period $`T=T_1+T_2`$, which is plotted in Fig. 2(a). The equations describing this traveling potential ratchet can be recast into the form of a tilting ratchet, which ordinarily would be implemented by applying an oscillatory external force to objects on an otherwise fixed landscape. The appropriate coordinate transformation, $`y(t)=x(t)f(t)`$ Reimann (2002), yields $$\eta \dot{y}(t)=V^{}(y(t))+F(t)+\xi (t),$$ (3) where $`F(t)=\eta \dot{f}(t)`$ is the effective driving force. Because $`f(t)`$ has a vanishing mean, the average velocity of the original problem is the same as that of the transformed tilting ratchet $`\dot{x}=\dot{y}`$, where the angle brackets imply both an ensemble average and an average over a period $`T`$. Reimann has demonstrated Reimann (2001, 2002) that a steady-state flux, $`\dot{y}0`$, develops in any tilting ratchet that breaks both spatiotemporal symmetry, $$V(y)=V(y)\text{, and }F(t)=F(t+T/2),$$ (4) and also spatiotemporal supersymmetry, $$V(y)=V(y+L/2)\text{, and }F(t+\mathrm{\Delta }t)=F(t),$$ (5) for any $`\mathrm{\Delta }t`$. No flux results if either of Eqs. (4) or (5) is satisfied. The optical trapping potential depicted in Fig. 1 is symmetric but not supersymmetric. Provided that $`F(t)`$ violates the symmetry condition in Eq. (4), the ratchet must induce directed motion. Although $`F(t)`$ is supersymmetric, as can be seen in Fig. 2(b), it is symmetric only when $`T_1=T_2`$. Consequently, we expect a particle current for $`T_1T_2`$. The zero crossing at $`T_1=T_2`$ furthermore portends flux reversal on either side of the equality. We calculate the steady-state velocity for this system by solving the master equation associated with Eq. (1) Risken (1989); Lee et al. (2005). The probability for a driven Brownian particle to drift from position $`x_0`$ to within $`dx`$ of position $`x`$ during the interval $`t`$, is given by the propagator $$P(x,t|x_0,0)dx=e^{^tL(x,t^{})𝑑t^{}}\delta (xx_0)dx,$$ (6) where the Liouville operator is $$L(x,t)=D\left(\frac{^2}{x^2}+\beta \frac{}{x}V^{}(x,t)\right),$$ (7) and where $`\beta ^1`$ is the thermal energy scale Risken (1989). The steady-state particle distribution $`\rho (x)`$ is an eigenstate of the master equation $$\rho (x)=P(x,T|x_0,0)\rho (x_0)𝑑x_0,$$ (8) and the associated steady-state flux is Lee et al. (2005) $$v=\frac{xx_0}{T}\rho (x_0)P(x,T|x_0,0)𝑑x𝑑x_0.$$ (9) The natural length scale in this problem is $`L`$, the inter-trap spacing in either state. The natural time scale, $`\tau =L^2/(2D)`$, is the time required for particles of diffusion constant $`D`$ to diffuse this distance. Figure 3 shows how $`v`$ varies with $`T_1/T_2`$ for various values of $`T/\tau `$ for experimentally accessible values of $`V_0`$, $`\sigma `$, and $`L`$. As anticipated, the net drift vanishes for $`T_1=T_2`$. Less obviously, the induced flux is directed from each well in the longer-duration state toward the nearest well in the short-lived state. The flux falls off as $`1/T`$ in the limit of large $`T`$ because the particles spend increasingly much of their time localized in traps. It also diminishes for short $`T`$ because the particles cannot keep up with the landscape’s evolution. In between, the range of fluxes can be tuned with $`T`$. We implemented this model for a sample of 1.53 $`\mu \mathrm{m}`$ diameter colloidal silica spheres (Bangs Laboratories, lot number 5328) dispersed in water, using potential energy landscapes created from arrays of holographic optical traps Dufresne and Grier (1998); Dufresne et al. (2001); Curtis et al. (2002); Lee et al. (2005). The sample was enclosed in a hermetically sealed glass chamber roughly 40 $`\mu \mathrm{m}`$ thick created by bonding the edges of a coverslip to a microscope slide, and was allowed to equilibrate to room temperature ($`21\pm 1^{}\mathrm{C}`$) on the stage of a Zeiss S100 2TV Axiovert inverted optical microscope. A $`100\times `$ NA 1.4 oil immersion SPlan Apo objective lens was used to focus the optical tweezer array into the sample and to image the spheres, whose motions were captured with an NEC TI 324A low noise monochrome CCD camera. The micrograph in Fig. 4(a) shows the focused light from a $`5\times 20`$ array of optical traps formed by a phase hologram projected with a Hamamatsu X7550 spatial light modulator Igasaki et al. (1999). The tweezers are arranged in twenty-trap manifolds $`37\mu \mathrm{m}`$ long separated by $`L=5.2\mu \mathrm{m}`$. Each trap is powered by an estimated $`2.5\pm 0.4\mathrm{mW}`$ of laser light at 532 nm. The particles, which appear in the bright-field micrograph in Fig. 4(b), are twice as dense as water and sediment to the lower glass surface, where they diffuse freely in the plane with a measured diffusion coefficient of $`D=0.33\pm 0.03\mu \mathrm{m}^2/\mathrm{sec}`$. This establishes the characteristic time scale for the system of $`\tau =39.4\mathrm{sec}`$, which is quite reasonable for digital video microscopy studies. Out-of-plane fluctuations were minimized by focusing the traps at the spheres’ equilibrium height above the wall Behrens et al. (2003). We projected two-state cycles of optical trapping patterns in which the manifolds in Fig. 4(a) were alternately displaced in the spheres’ equilibrium plane by $`L/3`$, with the duration of the first state fixed at $`T_1=3\mathrm{sec}`$ and $`T_2`$ ranging from 0.8 sec to 14.7 sec. To measure the flux induced by this cycling potential energy landscape for one value of $`T_2`$, we first gathered roughly two dozen particles in the middle row of traps in state 1, as shown in Fig. 4(b), and then projected up to one hundred periods of two-state cycles. The particles’ motions were recorded as uncompressed digital video streams for analysis Crocker and Grier (1996). Their time-resolved trajectories then were averaged over the transverse direction into the probability density, $`\rho (x,t)\mathrm{\Delta }x`$, for finding particles within $`\mathrm{\Delta }x=0.13\mu \mathrm{m}`$ of position $`x`$ after time $`t`$. We also tracked particles outside the trapping pattern to monitor their diffusion coefficients and to ensure the absence of drifts in the supporting fluid. Starting from this well-controlled initial condition resolves any uncertainties arising from the evolution of nominally random initial conditions Lee et al. (2005). Figures 4(c) and (d) show the spatially-resolved time evolution of $`\rho (x,t)`$ for $`T_2=0.8\mathrm{sec}<T_1`$ and $`T_2=8.6\mathrm{sec}>T_1`$. In both cases, the particles spend most of their time localized in traps, visible here as bright stripes, occasionally using the shorter-lived traps as springboards to neighboring wells in the longer-lived state. The mean particle position $`x(t)=x\rho (x,t)𝑑x`$ advances as the particles make these jumps, with the associated results plotted in Fig. 4(e). The speed with which an initially localized state, $`\rho (x,0)\delta (x)`$, advances differs from the steady-state speed plotted, in Fig. 3, but still can be calculated as the first moment of the propagator, $$x(t)=yP(y,t|0,0)𝑑y.$$ (10) Numerical analysis reveals a nearly constant mean speed that agrees quite closely with the steady-state speed from Eq. (9). Fitting traces such as those in Fig. 4(e) to linear trends provides estimates for the ratchet-induced flux, which are plotted in Fig. 4(f). The solid curve in Fig. 4(f) shows excellent agreement with predictions of Eq. (10) for $`\beta V_0=2.75\pm 0.5`$ and $`\sigma =0.65\pm 0.05\mu \mathrm{m}`$. Our implementation of the two-state ratchet involves updating the optical intensity pattern to translate the physical landscape. However, the same principles can be applied to systems in which the landscape remains fixed and the object undergoes cyclic transitions between two states. Figure 5 depicts a model for an active two-state walker on a fixed physical landscape that is inspired by the biologically relevant transport of single myosin head groups along actin filaments Kitamura et al. (1999). The walker consists of a head group that interacts with localized potential energy wells periodically distributed on the landscape. It also is attached to a lever arm that uses an external energy source to translate the head group by a distance somewhat smaller than the inter-well separation. The other end of the lever arm is connected to the payload, whose viscous drag would provide the leverage necessary to translate the head group between the extended and retracted states. Switching between the walker’s two states is equivalent to the two-state translation of the potential energy landscape in our experiments, and thus would have the effect of translating the walker in the direction of the shorter-lived state. A similar model in which a two-state walker traverses a spatially asymmetric potential energy landscape yields deterministic motion at higher efficiency than the present model Jülicher et al. (1997). It does not, however, allow for reversibility. The length of the lever arm and the diffusivity of the motor’s body and payload determine the ratio $`T/\tau `$ and thus the motor’s efficiency. The two-state ratchet’s direction does not depend on $`T/\tau `$, however, even under heavy loading. This differs from the three-state ratchet Lee et al. (2005), in which $`T/\tau `$ also controls the direction of motion. This protocol could be used in the design of mesoscopic motors based on synthetic macromolecules or microelectromechanical systems (MEMS). We are grateful for Mark Ofitserov’s many technical contributions. This work was supported by the National Science Foundation through Grant Number DBI-0233971 and Grant Number DMR-0304906. S.L. acknowledges support from a Kessler Family Foundation Fellowship
warning/0506/hep-th0506100.html
ar5iv
text
# 1 Introduction ## 1 Introduction We follow the analysis of Dimock , concerning the construction of the quantum Hilbert space of the Bosonic string and of the superstring the purpose being of presenting the facts as elementary as possible and in the same time rigorous. To be able to do that we start with a very simple method of constructing representations of the Virasoro and Kac-Moody algebras acting in Bosonic and Fermionic Fock spaces. We present a different way of computing things based on Wick theorem. Then we remind the main ingredients of the light-cone formalism and we prove the Poincaré invariance of the string and superstring systems. Most of the results obtained in this paper are known in the standard literature , , , , but we offer some new simple proofs. There are various attempts to clarify the main mathematical aspects of this topics (see the references) but we are closest to the spirit of , , and ,. To establish the equivalence between the light-cone and the covariant formalism one needs the so-called DDF operators. Usually DDF operators are introduced using formal series which are elements of the so-called vertex algebras . We present here an elementary derivation in the Bosonic case without using vertex algebras. Next we show that the covariant construction is equivalent (in the sense of group representation theory) with the light-cone construction; we are using the Hilbert space fiber-bundle formalism . Finally we give an elementary proof of the BRST quantization procedure for all the string models considered previously. In particular we are able to find very explicit formulas for the cohomology of the BRST operator. ## 2 Quadratic Hamiltonians in Fock spaces ### 2.1 Bose systems of oscillators We consider the Hilbert space $``$ with the scalar product $`<,>`$ generated by $`N`$ Bose oscillators; the creation and annihilation operators are $`a_m,a_m^+,m=1,\mathrm{},N`$ and verify $$[a_m,a_n^+]=\delta _{mn}I[a_m,a_n]=0,[a_m^+,a_n^+]=0,m,n=1,\mathrm{},N.$$ (2.1) If $`\mathrm{\Omega }`$ is the vacuum state we have $`a_m\mathrm{\Omega }=0,m>0.`$ As usual it is more convenient to introduce the operators $`\alpha _m,m=\{\pm 1,\mathrm{},\pm N\}`$ according to: $$\alpha _m=\sqrt{m}a_m,m>0\alpha _m=a_m^+,m<0$$ (2.2) Then the canonical commutation relation from above can be compactly written as follows: $$[\alpha _m,\alpha _n]=m\delta _{m+n}I,m,n0$$ (2.3) where $`\delta _m=\delta _{m,0}`$ and we also have $`\alpha _m\mathrm{\Omega }=0,m>0`$ $`\alpha _m^+=\alpha _mm.`$ (2.4) To apply Wick theorem we will also need the $`2`$-point function; we easily derive $$<\mathrm{\Omega },\alpha _m\alpha _n\mathrm{\Omega }>=\theta (m)m\delta _{m+n},m,n0$$ (2.5) where $`\theta `$ is the usual Heaviside function. The main result is the following ###### Theorem 2.1 Let us consider operators of the form $$H(A)\frac{1}{2}A_{mn}:\alpha _m\alpha _n:$$ (2.6) where $`A`$ is a symmetric matrix $`A^T=A`$ and the double dots give the Wick ordering. Then: $$[H(A),H(B)]=H([A,B])+\omega _\alpha (A,B)I$$ (2.7) where the commutator $`[A,B]=ABBA`$ is computed using the following matrix product $$(AB)_{pq}\underset{m0}{}mA_{pm}B_{m,q}$$ (2.8) and we have defined $$\omega _\alpha (A,B)\frac{1}{2}\underset{m,n>0}{}mnA_{mn}B_{n,m}(AB).$$ (2.9) Proof: Is elementary and relies on computing the expression $`H(A)H(B)`$ using Wick theorem: $`H(A)H(B)=:H(A)H(B):`$ $`+{\displaystyle \frac{1}{4}}A_{mn}B_{pq}[<\mathrm{\Omega },\alpha _m\alpha _p\mathrm{\Omega }>:\alpha _n\alpha _q:+(mn)+(pq)+(mn,pq)]`$ $`+<\mathrm{\Omega },\alpha _m\alpha _p\mathrm{\Omega }><\mathrm{\Omega },\alpha _n\alpha _q\mathrm{\Omega }>I+<\mathrm{\Omega },\alpha _m\alpha _q\mathrm{\Omega }><\mathrm{\Omega },\alpha _n\alpha _p\mathrm{\Omega }>I]`$ (2.10) If we use the $`2`$-point function we easily arrive at the formula from the statement. $`\mathrm{}`$ Now we extend the previous result to the case when we have an infinite number of oscillators. We consider that $`m^{}`$ and the matrix $`A`$ is semi-finite i.e. there exists $`N>0`$ such that $$A_{mn}=0\mathrm{for}|m+n|>N.$$ (2.11) We note that if $`A`$ and $`B`$ are semi-finite then $`AB`$ is also semi-finite. We will need the algebraic Fock space which is the subspace $`𝒟_0`$ with finite number of particles. The elements of $`𝒟_0`$ are, by definition, finite linear combinations of vectors of the type $`a_{m_1}^+\mathrm{}a_{m_k}^+\mathrm{\Omega };`$ the subspace $`𝒟_0`$ is dense in $``$. Then one can prove easily that the operator $`H(A)`$ is well defined on $`𝒟_0`$, leaves $`𝒟_0`$ invariant and formula (2.7) remains true in $`𝒟_0`$. We will need an extension of this result namely we want to consider the case when the index $`m`$ takes the value $`m=0`$ also i.e. $`m`$ and we preserve the commutation relation (2.3). We note that the relation (2.5) is not valid if one (or both) of the indices are null so the previous proof does not work. It can be proved, however, directly that the statement of the theorem remains true if we extend accordingly the definition of the matrix product including the value $`0`$ also i.e in (2.8) we leave aside the restriction $`m=0`$. In general, the Hilbert space of this case will not be entirely of Fock type: the operators $`\alpha _mm0`$ will live in a Fock space tensored with another Hilbert space where live the operators $`\alpha _0`$. ### 2.2 A Systems of Fermi Oscillators We consider the Hilbert space $``$ with the scalar product $`<,>`$ generated by $`N`$ Fermi oscillators; the creation and annihilation operators are $`b_m,b_m^+,m=1,\mathrm{},N`$ and verify $$\{b_m,b_n^+\}=\delta _{mn}I\{b_m,b_n\}=0,\{b_m^+,b_n^+\}=0,m,n0.$$ (2.12) If $`\mathrm{\Omega }`$ is the vacuum state we have $`b_m\mathrm{\Omega }=0,m>0.`$ As above it is more convenient to introduce the operators $`b_m,m=\{\pm 1,\mathrm{},\pm N\}`$ according to: $$b_m=b_m,m>0b_m=b_m^+,m<0$$ (2.13) and the canonical anti-commutation relation from above can be compactly rewritten as follows: $$\{b_m,b_n\}=\delta _{m+n}I,m,n0.$$ (2.14) We also have $`b_m\mathrm{\Omega }=0,m>0`$ $`b_m^+=b_m.`$ (2.15) The $`2`$-point function is in this case: $$<\mathrm{\Omega },b_mb_n\mathrm{\Omega }>=\theta (m)\delta _{m+n},m,n0.$$ (2.16) The main result is the following ###### Theorem 2.2 Let us consider operators of the form $$H(A)\frac{1}{2}A_{mn}:b_mb_n:$$ (2.17) where $`A`$ is a antisymmetric matrix $`A^T=A`$ and the double dots give the Wick ordering. Then: $$[H(A),H(B)]=H([A,B])+\omega _b(A,B)I$$ (2.18) where the commutator $`[A,B]=ABBA`$ is computed using the following matrix product $$(AB)_{pq}\underset{m0}{}A_{pm}B_{m,q}$$ (2.19) and we have defined $$\omega _b(A,B)\frac{1}{2}\underset{m,n>0}{}A_{mn}B_{n,m}(AB).$$ (2.20) The proof is similar to the preceding theorem. The previous result can be extended to the case when we have an infinite number of oscillators i.e. $`m^{}`$ and the matrix $`A`$ is semi-finite: the operator $`H(A)`$ is well defined on the corresponding algebraic Fock space $`𝒟_0`$, leaves invariant this subspace and the previous theorem remains true. ### 2.3 Another System of Fermi Oscillators We extend the previous results to the case when the value $`m=0`$ is allowed i.e. the Hilbert space $``$ is generated by the operators $`d_m,m=N,\mathrm{},N`$ and verify $`\{d_m,d_n\}=\delta _{mn}Im,n`$ $`d_m^{}=d_mm`$ $`d_m\mathrm{\Omega }=0m>0`$ (2.21) where $`\mathrm{\Omega }`$ is the vacuum state. One can realize this construction if one takes $`=^s𝒞`$ where $``$ is the Fock space from the preceding Section, $`𝒞`$ is the Clifford algebra generated by the element $`b_0`$ verifying $`b_0^{}=b_0b_0^2=1/2`$ and the skew tensor product $`^s`$ is chosen such that the operators $$d_nb_n^sI_2m0d_0I_1^sb_0$$ (2.22) verify (2.21). Another more explicit construction is to consider the Hilbert space is generated by the creation and annihilation operators $`b_m,b_m^+,m=0,\mathrm{},N`$ such that we have $`b_m\mathrm{\Omega }=0,m0`$ and to define the operators $`d_m,m=\{N,\mathrm{},N\}`$ according to: $`d_m=\{\begin{array}{cc}b_m,\hfill & \text{for m }>\text{ 0}\hfill \\ \frac{1}{\sqrt{2}}(b_0+b_0^+),\hfill & \text{for m = 0}\hfill \\ b_m^+,\hfill & \text{for m }<\text{ 0}\hfill \end{array}`$ (2.23) The $`2`$-point function is in this case: $$<\mathrm{\Omega },d_md_n\mathrm{\Omega }>=\theta _+(m)\delta _{m+n},m,n$$ (2.24) where we have introduced the modified Heaviside function $`\theta _+(m)=\{\begin{array}{cc}1,\hfill & \text{for m }>\text{ 0}\hfill \\ \frac{1}{2},\hfill & \text{for m = 0}\hfill \\ 0,\hfill & \text{for m }<\text{ 0}\hfill \end{array}.`$ (2.25) It follows that the main result is similar to the previous one: ###### Theorem 2.3 Let us consider operators of the form $$H(A)\frac{1}{2}A_{mn}:d_md_n:$$ (2.26) where $`A`$ is a antisymmetric matrix $`A^T=A`$ and the double dots give the Wick ordering. Then: $$[H(A),H(B)]=H([A,B])+\omega _d(A,B)I$$ (2.27) where the commutator $`[A,B]=ABBA`$ is computed using the following matrix product $$(AB)_{pq}\underset{m=N}{\overset{N}{}}A_{pm}B_{m,q}$$ (2.28) and we have defined $$\omega _d(A,B)\frac{1}{2}\underset{m,n0}{}\theta _+(m)\theta _+(n)A_{mn}B_{n,m}(AB).$$ (2.29) The proof is similar to the preceding Section. The previous result can be extended to the case when we have an infinite number of oscillators i.e. $`m`$ and the matrix $`A`$ is semi-finite: the operator $`H(A)`$ is well defined on the corresponding algebraic Fock space $`𝒟_0`$, leaves invariant this subspace and the previous theorem remains true. ###### Remark 2.4 It follows easily that the expressions $`\omega _\alpha ,\omega _b,\omega _d`$ are $`2`$-cocyles. They are quantum obstructions (or anomalies) because they do not appear if we work in classical field theory replacing the commutators by Poisson brackets. ## 3 Virasoro Algebras in Fock Spaces We have constructed some Fock spaces for which we have a nice commutation relation of the bilinear operators. In all these cases we will be able to construct representations of the Virasoro algebra taking convenient expressions for the matrix $`A`$. We give the details corresponding to the structure of the closed strings. ### 3.1 Bose Case ###### Theorem 3.1 In the conditions of theorem 2.9 the operators given by the formulas $$L_m\frac{1}{2}\underset{n0,m}{}:\alpha _{mn}\alpha _n:$$ (3.1) are well defined on the algebraic Fock subspace, leave invariant this subspace and verify the following relations: $`[L_m,L_n]=(mn)L_{m+n}+{\displaystyle \frac{m(m^21)}{12}}\delta _{m+n}I`$ $`L_m^+=L_mm`$ $`L_0\mathrm{\Omega }=0.`$ (3.2) Proof: We consider the matrices $`A_m`$ given by $`(A_m)_{pq}\delta _{p+qm}`$ and we are in the conditions of theorem 2.9: the matrices $`A_m`$ are symmetric and semi-finite. It remains to prove that $`[A_m,A_n]=(mn)A_{m+n}`$ and to compute the $`2`$-cocycle $`\omega _\alpha (A_m,A_n)=\frac{m(m^21)}{12}\delta _{m+n}`$ and we obtain the commutation relation from the statement. $`\mathrm{}`$ One can express everything in terms of the original creation and annihilation operators $`a_m^\mathrm{\#}`$; if we use the holomorphic representation for the harmonic oscillator operators $`a_m^+=z_ma_m=\frac{}{z_m}`$ we obtain the formula (7.2.10) from . It is important that we can extend the previous results to the case when $`\alpha _00`$ (see the end of Subsection 2.1.) To preserve (2.3) we impose $$[\alpha _0,\alpha _m]=0m^{}$$ (3.3) and we keep the relation (3.1) without the restrictions $`n0,m`$; explicitly: $$L_m=\mathrm{}+\alpha _m\alpha _0m0,L_0=\mathrm{}+\frac{1}{2}\alpha _0^2$$ (3.4) where by $`\mathrm{}`$ we mean the expressions from the preceding theorem corresponding to $`\alpha _00`$. Eventually we have to consider a larger Hilbert space containing as a subspace the Fock space generated by the operators $`\alpha _nn0`$. By direct computations we can prove that the statement of the theorem remains true; also we have $$[L_m,\alpha _n]=n\alpha _{m+n}.$$ (3.5) In the following we will use only the case when $`\alpha _00`$. ### 3.2 First Fermi Case We have a similar result for the Fermi operators of type $`b`$: we will consider that these operators are $`b_r`$ indexed by $`r\frac{1}{2}+`$ and they verify: $`\{b_r,b_s\}=\delta _{r+s}I,r,s{\displaystyle \frac{1}{2}}+`$ $`b_r\mathrm{\Omega }=0,r>0`$ $`b_r^+=b_r.`$ (3.6) Then: ###### Theorem 3.2 In the conditions of theorem 2.20 the operators given by the formulas $$L_m\frac{1}{2}\underset{r1/2+}{}\left(r+\frac{m}{2}\right):b_rb_{m+r}:=\frac{1}{2}\underset{r1/2+}{}r:b_rb_{m+r}:$$ (3.7) are well defined on the algebraic Fock subspace, leave invariant this subspace and verify the following relations: $`[L_m,L_n]=(mn)L_{m+n}+{\displaystyle \frac{m(m^21)}{24}}\delta _{m+n}I.`$ $`[L_m,b_r]=\left(r+{\displaystyle \frac{m}{2}}\right)b_{m+r}`$ $`L_m^+=L_m`$ $`L_0\mathrm{\Omega }=0.`$ (3.8) Proof: We consider the matrices $`A_m`$ given by $`(A_m)_{rs}\frac{1}{2}(sr)\delta _{r+sm}`$ and we are in the conditions of theorem 2.20: the matrices $`A_m`$ are anti-symmetric and semi-finite. It remains to compute the $`2`$-cocycle $`\omega _b`$ to obtain the commutation relation from the statement. $`\mathrm{}`$ If we use the representation in terms of Grassmann variables $`b_r^+=\xi _rb_r=\frac{}{\xi _r}`$ for these operators we obtain the formulas from pg. 225. ### 3.3 Second Fermi Case Finally we have a similar result for the Fermi operators of type $`d`$. ###### Theorem 3.3 In the conditions of theorem 2.29 the operators given by the formulas $$L_m\frac{1}{2}\underset{n}{}\left(n+\frac{m}{2}\right):d_nd_{m+n}:=\frac{1}{2}\underset{n}{}n:d_nd_{m+n}:$$ (3.9) are well defined on the algebraic Fock subspace, leave invariant this subspace and verify the following relations: $`[L_m,L_n]=(mn)L_{m+n}+{\displaystyle \frac{m(m^2+2)}{24}}\delta _{m+n}I.`$ $`[L_m,d_n]=\left(n+{\displaystyle \frac{m}{2}}\right)b_{m+n}`$ $`L_m^+=L_m`$ $`L_0\mathrm{\Omega }=0.`$ (3.10) Proof: We consider the matrices $`A_m`$ given by $`(A_m)_{pq}\frac{1}{2}(qp)\delta _{p+qm}`$ and we are in the conditions of theorem 2.29: the matrices $`A_m`$ are anti-symmetric and semi-finite. It remains to compute the $`2`$-cocycle $`\omega _d`$ to obtain the commutation relation from the statement. $`\mathrm{}`$ We observe that in the commutation relation of the preceding theorem the expression of the cocycle is different from the usual form $`c\frac{m(m^21)}{12};`$ we can fix this inconvenience immediately if we define: $$\stackrel{~}{L}_mL_mm0\stackrel{~}{L}_0L_0+\frac{1}{16}I;$$ (3.11) we obtain in this case: $$[\stackrel{~}{L}_m,\stackrel{~}{L}_n]=(mn)\stackrel{~}{L}_{m+n}+\frac{m(m^21)}{24}\delta _{m+n}I.$$ (3.12) and $$\stackrel{~}{L}_0\mathrm{\Omega }=\frac{1}{16}I.$$ (3.13) ### 3.4 Multi-Dimensional Cases In the preceding Subsections we have obtained three representations of the Virasoro algebra corresponding to $`(c,h)=(1,0),(\frac{1}{2},0),(\frac{1}{2},\frac{1}{16}).`$ The previous results can be easily extended to a more general case. Let $`\eta ^{jk}=\eta _{jk},j,k=1,\mathrm{},D`$ be a diagonal matrix with the diagonal elements $`ϵ_1,\mathrm{},ϵ_D=\pm 1.`$ In the Bose case we can consider that we have the family of operators: $`\alpha _m^j,m,j=1,\mathrm{},D`$ acting in the Hilbert space $`^{(\alpha )}`$ such that: $`[\alpha _m^j,\alpha _n^k]=m\eta _{jk}\delta _{m+n}I,m,n`$ $`\alpha _m^j\mathrm{\Omega }=0,m>0`$ $`(\alpha _m^j)^+=\alpha _m^jm.`$ (3.14) We can define $$L_m^{(\alpha )}\frac{1}{2}\underset{n}{}\eta _{jk}:\alpha _{mn}^j\alpha _n^k:$$ (3.15) and we have: $$[L_m^{(\alpha )},L_n^{(\alpha )}]=(mn)L_{m+n}^{(\alpha )}+D\frac{m(m^21)}{12}\delta _{m+n}I.$$ (3.16) In the first Fermi case we have the operators: $`b_r^j,r\frac{1}{2}+,j=1,\mathrm{},D`$ acting in the Hilbert space $`^{(b)}`$ such that: $`\{b_r^j,b_s^k\}=\eta _{jk}\delta _{r+s}I,r,s`$ $`b_r^j\mathrm{\Omega }=0,r>0`$ $`(b_r^j)^+=b_r^jr.`$ (3.17) We define $$L_m^{(b)}\frac{1}{2}\underset{r1/2+}{}\left(r+\frac{m}{2}\right)\eta _{jk}:b_r^jb_{m+r}^k:$$ (3.18) are well defined and verify the following relations: $$[L_m^{(b)},L_n^{(b)}]=(mn)L_{m+n}^{(b)}+D\frac{m(m^21)}{24}\delta _{m+n}I.$$ (3.19) Finally, in the second Fermi case we have the operators $`d_m^j,m,j=1,\mathrm{},D`$ acting in the Hilbert space $`^d`$ such that $`\{d_m^j,d_n^k\}=\eta _{jk}\delta _{m+n}I`$ $`d_m^j\mathrm{\Omega }=0,m>0`$ $`(d_m^j)^+=d_m^jm.`$ (3.20) We can define $$L_m^{(d)}\frac{1}{2}\underset{n}{}\left(n+\frac{m}{2}\right)\eta _{jk}:d_n^jd_{m+n}^k:$$ (3.21) and we have the following relations: $$[L_m^{(d)},L_n^{(d)}]=(mn)L_{m+n}^{(d)}+D\frac{m(m^2+2)}{24}\delta _{m+n}I.$$ (3.22) We redefine $$\stackrel{~}{L}_m^{(d)}L_m^{(d)}m0\stackrel{~}{L}_0^{(d)}L_0^{(d)}+\frac{D}{16}I;$$ (3.23) we obtain in this case: $$[\stackrel{~}{L}_m^{(d)},\stackrel{~}{L}_n^{(d)}]=(mn)\stackrel{~}{L}_{m+n}^{(d)}+D\frac{m(m^21)}{24}\delta _{m+n}I.$$ (3.24) $$\stackrel{~}{L}_0^{(d)}\mathrm{\Omega }=\frac{D}{16}I.$$ (3.25) In all these cases the $`2`$-cocycle gets multiplied by $`D`$. The Hilbert space has a positively defined scalar product only in the case $`ϵ_1=\mathrm{}=ϵ_D=1.`$ We can combine the Bose and Fermi cases as follows. We consider the Hilbert spaces $`^{(NS)}^{(\alpha )}^{(b)}`$ and $`^{(R)}^{(\alpha )}^{(d)}`$ respectively; the Virasoro operators are $`L_m^{(NS)}L^{(\alpha )}I_2+I_1L^{(b)}`$ $`L_m^{(R)}L^{(\alpha )}I_2+I_1\stackrel{~}{L}^{(d)}`$ (3.26) and we have in both cases: $$[L_m^{(NS,R)},L_n^{(NS,R)}]=(mn)L_{m+n}^{(NS,R)}+D\frac{m(m^21)}{8}\delta _{m+n}I.$$ (3.27) These two constructions are called Neveau-Schwartz (and Ramond) respectively. In these cases one can extend the Virasoro algebra to a super-Virasoro algebra . We conclude this Subsection with some simple propositions. First we have a natural representation of the rotation group in the Fock space: ###### Proposition 3.4 Suppose that the signature of $`\eta `$ is $`(r,s)`$; then we can define in the corresponding Hilbert spaces a representation of the Lie algebra $`so(r,s)`$ according to: $`J^{(\alpha )jk}i{\displaystyle \underset{m>0}{}}{\displaystyle \frac{1}{m}}\alpha _m^j\alpha _m^k(jk)`$ $`J^{(b)jk}i{\displaystyle \underset{r>0}{}}b_r^jb_r^k(jk)`$ $`J^{(d)jk}i{\displaystyle \underset{m>0}{}}d_m^jd_m^k(jk)`$ (3.28) respectively. Indeed, we can obtain directly from the (anti)commutation relations in all the cases: $$[J^{kl},J^{pq}]=i(\eta ^{lp}J^{jq}+\eta ^{jq}J^{lp}\eta ^{kp}J^{lq}+\eta ^{lq}J^{kp}).$$ (3.29) We also note that the Virasoro operators are rotational invariant: in all cases $$[J^{kl},L_m]=0.$$ (3.30) Next, we have a proposition which will be important for proving the Poincaré invariance: ###### Proposition 3.5 If $`\mathrm{\Psi }𝒟_0`$ is an arbitrary vector from the algebraic Hilbert space then we have in all cases $$L_m\mathrm{\Psi }=0$$ (3.31) for sufficiently large $`m>0`$. Proof: We consider only the one-dimensional Bose case; the other cases are similar. If $`\mathrm{\Psi }`$ is a vector in the algebraic Fock space it is clear that we have $$\alpha _m\mathrm{\Psi }=0$$ (3.32) for $`m`$ sufficiently large. This implies immediately that all the sums in (3.1) are finite (it is better to re-express everything in terms of the original creation and annihilation operators). It is clear that $`_{n>0}\mathrm{}a_n^+a_{m+n}\mathrm{\Psi }=0`$ for sufficiently large $`m`$ because of (3.32). Also $`_{n=1}^{m1}\mathrm{}a_na_{mn}\mathrm{\Psi }=0`$ for sufficiently large $`m`$; indeed, all the indices in the preceding sum are larger than $`\frac{m}{2}`$ so again we can apply (3.32). $`\mathrm{}`$ It is known that any $`2`$-cocycle of the Virasoro algebra is cohomologous to a standard form $`c\frac{m(m^21)}{12};`$ however, we can always add a trivial cocycle. ###### Proposition 3.6 Suppose that we have $$[L_m,L_n]=(mn)L_{m+n}+c\frac{m(m^21)}{12}\delta _{m+n}I$$ (3.33) and we redefine $$L_m(a)L_mm0L_0(a)L_0aI;a.$$ (3.34) Then we have: $$[L_m(a),L_n(a)]=(mn)L_{m+n}(a)+\left[c\frac{m(m^21)}{12}+2am\right]\delta _{m+n}I.$$ (3.35) ## 4 Kac-Moody Algebras in Fock Spaces We use here the techniques from Section 2. We suppose that we have $`t_A,A=1,\mathrm{},r`$ a $`N`$-dimensional representation of the Lie algebra $`𝔤`$: $$[t_A,t_B]=f_{ABC}t_C;$$ (4.1) here $`f_{ABC}`$ are the structure constants of $`𝔤`$ and we need also the Killing-Cartan form of the representation $`t`$: $$g_{AB}Tr(t_At_B).$$ (4.2) We will suppose that the matrices $`t_A`$ are antisymmetric: $$t_A^T=t_A.$$ (4.3) Finally, we will need the contra-gradient representation $$\stackrel{~}{t}_A=t_A^+=\overline{t}_AA.$$ (4.4) We will construct some representations of the Kac-Moody algebras acting in $`^{(b)}`$ and $`^{(d)}`$ respectively. ### 4.1 Neveau-Schwartz Case In this case ###### Theorem 4.1 In the conditions of theorem 2.20 the operators given by the formulas $$K_m^A\frac{1}{2}(t_A)_{jk}\underset{r1/2+}{}:b_r^jb_{m+r}^k:$$ (4.5) are well defined on the algebraic Fock subspace, leave invariant this subspace and verify the following relation: $$[K_m^A,K_n^B]=f_{ABC}K_{m+n}^C+\frac{1}{2}mg_{AB}\delta _{m+n}I.$$ (4.6) We have in this case also the Hermiticity property: $$(K_m^A)^+=\stackrel{~}{K}_m^A$$ (4.7) where $`\stackrel{~}{K}_m^A`$ is associated to the representation $`\stackrel{~}{t}_A.`$ The following commutation relation are true: $$[K_m^A,L_n]=mK_{m+n}^A.$$ (4.8) If $`\mathrm{\Psi }𝒟_0`$ then we have $$K_m^A\mathrm{\Psi }=0$$ (4.9) for $`m>0`$ large enough. Proof: We consider the matrices $`A_m^A`$ given by $`A_m^A\stackrel{~}{A}_mt_A`$ where $`(\stackrel{~}{A}_m)_{rs}\delta _{r+sm}`$ and we are in the conditions of theorem 2.20: the matrices $`A_m^A`$ are anti-symmetric (because $`\stackrel{~}{A}_m`$ are symmetric and $`t_A`$ are anti-symmetric) and semi-finite. We can easily prove that $`[A_m^A,A_n^B]=f_{ABC}A_{m+n}^C;`$ it remains to compute the $`2`$-cocycle $`\omega _b`$ to obtain the commutation relation (4.6). The relation (4.8) follows from the same theorem 2.20 if we take $`AA_m^A,BA_nI_2;`$ in this case the cocycle $`\omega _b`$ is null because $`Tr(t_A)=0.`$ The relation (4.9) is proved as Proposition 3.5. $`\mathrm{}`$ We have an important particular case: ###### Proposition 4.2 Let us consider the case $`𝔤=so(D)`$ with the fundamental representation $$(E_{jk})_{pq}=\delta _{jp}\delta _{kq}\delta _{jq}\delta _{kp}.$$ (4.10) Then the associated operators, denoted by $`K_m^{jk}`$ verify the following relations: $$[K_m^{jk},K_n^{pq}]=\delta _{kp}K_{m+n}^{jq}\delta _{kq}K_{m+n}^{jp}+\delta _{jq}K_{m+n}^{kp}\delta _{jp}K_{m+n}^{kq}+m(\delta _{kp}\delta _{jq}\delta _{kq}\delta _{jp})\delta _{m+n}I$$ (4.11) $$(K_m^{jk})^+=K_m^{jk}$$ (4.12) Proof: We apply theorem 2.20; in our case we have by direct computation $$[E_{jk},E_{pq}]=\delta _{kp}E_{jq}\delta _{kq}E_{jp}+\delta _{jq}E_{kp}\delta _{jp}E_{kq}$$ (4.13) and the Killing-Cartan form $$g_{jk,pq}=2(\delta _{kp}\delta _{jq}\delta _{kq}\delta _{jp}).$$ (4.14) Then we apply the preceding theorem. $`\mathrm{}`$ From the commutation relation of the preceding Proposition we observe that \- see (3.28): $`J^{(b)jk}=iK_0^{jk}`$ so we have another proof that these operators give a representation of the group $`so(D).`$ Finally we give the following technical result : ###### Proposition 4.3 In the preceding conditions, the following formula is true $`{\displaystyle \underset{m>0}{}}m[K_m^{jl}K_m^{kl}(jk)]+2{\displaystyle \underset{m>0}{}}(K_m^{jk}L_m+L_mK_m^{jk})+2K_0^{jk}L_0`$ $`={\displaystyle \frac{8D}{8}}{\displaystyle \underset{r>0}{}}(4r^21)[:b_r^jb_r^k:(jk)]+(12a)K_0^{jk};`$ (4.15) here $`L_m=L_m(a).`$ Proof: First we note that all the sums are in fact finite when applied on the algebraic Fock space. Then one substitutes the expressions for $`K`$’s and $`L`$’s and applies Wick theorem. The computation is tedious but straightforward. $`\mathrm{}`$ Let us note that we have a quantum anomaly in the right-hand side which vanishes for $`D=8,a=\frac{1}{2}.`$ ### 4.2 Ramond Case ###### Theorem 4.4 In the conditions of theorem 2.29 the operators given by the formulas $$K_m^A\frac{1}{2}(t_A)_{jk}\underset{n}{}:d_n^jd_{m+n}^k:$$ (4.16) are well defined on the algebraic Fock subspace, leave invariant this subspace and verify the same relations as in the preceding theorem. Proof: We consider the matrices $`A_m`$ given by $`A_m^A\stackrel{~}{A}_mt_A`$ where $`(A_m)_{pq}\delta _{p+qm}`$ and we are in the conditions of theorem 2.29. $`\mathrm{}`$ The technical result is in this case: ###### Proposition 4.5 In the preceding conditions, the following formula is true $`{\displaystyle \underset{m>0}{}}m[K_m^{jl}K_m^{kl}(jk)]+2{\displaystyle \underset{m>0}{}}(K_m^{jk}L_m+L_mK_m^{jk})+2K_0^{jk}L_0`$ $`={\displaystyle \frac{8D}{2}}{\displaystyle \underset{k>0}{}}k^2[:d_r^jd_r^k:(jk)]+({\displaystyle \frac{D}{8}}2a)K_0^{jk}.`$ (4.17) Again we have in the right-hand side a quantum anomaly which vanishes for $`D=8,a=\frac{1}{2}.`$ ###### Remark 4.6 We note that in the relations (4.15) and (4.17) the appearance of the constant $`a`$ is in fact spurious: if we express the Virasoro operators $`L_m`$ in terms of the operators $`\alpha ,b`$ or $`d`$ then the constant $`a`$ drops out. ## 5 Light-Cone Coordinates In this Section we remind the basic ingredients of the light-cone description of the representations of the Poincaré group. We consider this group in $`D`$ dimensions and we denote the indices by $`\mu ,\nu =0,1,\mathrm{},D1`$. Let $`\eta ^{\mu \nu }`$ be the Minkowski diagonal form with $`diag(\eta )=1,1,\mathrm{},1.`$ Then the algebra of the Poincaré group is generated by the basis elements $`P^\mu ,J^{\mu \nu }`$ where $`J^{\mu \nu }`$ is anti-symmetric. The Lie brackets are: $`[P^\mu ,P^\nu ]=0`$ $`[P^\mu ,J^{\nu \rho }]=i\eta ^{\mu \nu }P^\rho i\eta ^{\mu \rho }P^\nu `$ $`[J^{\mu \nu },J^{\rho \lambda }]=i\eta ^{\nu \rho }J^{\mu \lambda }i\eta ^{\mu \rho }J^{\nu \lambda }i\eta ^{\nu \lambda }J^{\mu \rho }+i\eta ^{\mu \lambda }J^{\nu \rho }.`$ (5.1) Let us change the basis $`(P^\mu ,J^{\mu \nu })(P^\pm ,P^j,J^+,J^{j\pm },J^{jk})`$ where $`j=1,\mathrm{},D2`$ and $`P^\pm {\displaystyle \frac{1}{\sqrt{2}}}(P^0\pm P^{D1})`$ $`J^{j\pm }{\displaystyle \frac{1}{\sqrt{2}}}(J^{j0}\pm J^{j,D1})`$ $`J^+J^{0,D1}.`$ (5.2) Then in the new basis the Lie brackets are: $`[P^{ϵ_1},P^{ϵ_2}]=0[P^ϵ,P^j]=0[P^j,P^k]=0`$ $`[P^ϵ,J^{jk}]=0[P^j,J^+]=0`$ $`[P^ϵ,J^+]=iϵP^ϵ`$ $`[P^{ϵ_1},J^{jϵ_2}]={\displaystyle \frac{i}{2}}(1ϵ_1ϵ_2)P^j`$ $`[P^j,J^{kl}]=i\delta _{jk}P^l+i\delta _{jl}P^k`$ $`[P^j,J^{kϵ}]=i\delta _{jk}P^ϵ`$ $`[J^+,J^{jk}]=0`$ $`[J^+,J^{jϵ}]=iϵJ^{jϵ}`$ $`[J^{jϵ_1},J^{kϵ_2}]={\displaystyle \frac{i}{2}}(1+ϵ_1ϵ_2)J^{jk}+{\displaystyle \frac{i}{2}}(ϵ_2ϵ_1)\delta _{jk}J^+`$ $`[J^{jϵ},J^{kl}]=i\delta _{jk}J^{lϵ}+i\delta _{jl}J^{kϵ}`$ $`[J^{kl},J^{pq}]=i\delta _{lp}J^{kq}+i\delta _{kp}J^{lq}+i\delta _{lq}J^{kp}i\delta _{kq}J^{lp};`$ (5.3) here $`j,k,l=1,\mathrm{},D2`$ and $`ϵ=\pm `$. We will need the representation of mass $`m>0`$ and spin $`0`$ in light-cone coordinates. The Hilbert space is $$^{[m,0]}^{[m]}L^2(^{D1},\frac{d𝐩}{2p^0})$$ (5.4) where $`d𝐩dp^1\mathrm{}dp^{D1},`$ $`p^0\sqrt{𝐩^2+m^2}`$ and $`𝐩^2=(𝐩^1)^2+\mathrm{}(𝐩^{D1})^2.`$ One can easily derive the expressions of the Lie algebra generators; they are: $`P^\mu =p^\mu `$ $`L^{jk}=i\left(p^j{\displaystyle \frac{}{p^k}}p^k{\displaystyle \frac{}{p^j}}\right)`$ $`K^jL^{j0}=ip^0{\displaystyle \frac{}{p^j}}`$ (5.5) where $`j,k=1,\mathrm{},D1`$. These operators are defined on $`𝒞^{\mathrm{}}(^{D1})`$ where they are essentially self-adjoint; one can easily verify that on this domain the relations (5.1) are valid. Now we perform a change of variables : $$(p^1,\mathrm{},p^{D1})(p^+,\stackrel{~}{p})$$ (5.6) where $$p^+\frac{1}{\sqrt{2}}(p^0+p^{D1})\stackrel{~}{p}=(p^1,\mathrm{},p^{D2});$$ (5.7) here, as before, we have denoted $`p^0\sqrt{𝐩^2+m^2}.`$ We will also denote $`\stackrel{~}{𝐩}^2=(𝐩^1)^2+\mathrm{}(𝐩^{D2})^2.`$ The converse of this change of variables is: $$p^{D1}=\frac{1}{\sqrt{2}}(p^+p^{})$$ (5.8) where we have introduced the notation $$p^{}=p^{}(p^+,\stackrel{~}{𝐩})\frac{\stackrel{~}{𝐩}^2+m^2}{2p^+}.$$ (5.9) Now we re-express the measure $`\frac{d𝐩}{2p^0}`$ in the new variables and easily get: $$\frac{d𝐩}{2p^0}=\frac{d\stackrel{~}{p}dp^+}{2p^+}.$$ (5.10) Let us note that for $`m>0`$ the variable $`p^+`$ takes values in the whole positive real axis so, in the new variables the Hilbert space is $$^{[m]}L^2(_+\times ^{D2},\frac{dp^+d\stackrel{~}{p}}{2p^+})$$ (5.11) where as usual $`_+(0,\mathrm{}).`$ The unitary transformation connecting the two representations is $`V:L^2(^{D1},\frac{d𝐩}{2p^0})L^2(_+\times ^{D2},\frac{dp^+d\stackrel{~}{p}}{2p^+})`$ is: $$(Vf)(p^+,\stackrel{~}{p})=f(\stackrel{~}{p},\frac{p^+p^{}}{\sqrt{2}});$$ (5.12) the inverse of this transformation can be easily provided. It is a straightforward exercise to derive the expressions of the Lie algebra generators in the light-cone coordinates: $`P^j=p^jP^+=p^+P^{}=p^{}{\displaystyle \frac{\stackrel{~}{𝐩}^2+m^2}{2p^+}}`$ $`L^{jk}=i\left(p^j{\displaystyle \frac{}{p^k}}p^k{\displaystyle \frac{}{p^j}}\right)L^{j+}=ip^+{\displaystyle \frac{}{p^j}}`$ $`L^j=i\left(p^{}{\displaystyle \frac{}{p^j}}+p^j{\displaystyle \frac{}{p^+}}\right)L^+=ip^+{\displaystyle \frac{}{p^+}}`$ (5.13) where $`j,k=1,\mathrm{},D2`$. These operators are defined on $`𝒟^{[m]}𝒞^{\mathrm{}}(_+\times ^{D2})`$ where they are essentially self-adjoint; one can easily verify that on this domain the relations (5.3) are valid. The extensions of the preceding formulas to the case of arbitrary $`m`$ (i.e. $`m=0`$ and $`m`$ purely imaginary) can be done using Lemma 3 of . We define for any $`j=1,\mathrm{},D1`$ expressions of the type $`p^+`$ namely $$p_j^+\frac{1}{\sqrt{2}}(p^0+p^j)$$ (5.14) and the chart $$V_j=\{(p_j^+,p^1,\mathrm{},p^{j1},p^{j+1},\mathrm{},p^{D1})_+\times ^{D2}\}.$$ (5.15) In every such chart there are no singularities and one can consider the corresponding measure defined as in (5.10). The reunion of all these charts is a cover of the whole mass shell and if we consider a partition of the unity subordinated to this partition then we can obtain the corresponding measure in light-cone coordinates. ## 6 Quantum Strings and Superstrings In this Section we use the preceding construction to give a proper mathematical definition for the string and superstring systems. ### 6.1 The Quantum Bosonic String We introduce the following construction. First we consider the system of Bose oscillators $`\alpha _m^j,m^{},j=1,\mathrm{},D2`$ (i.e. we exclude for the moment the value $`m=0`$). The Hilbert space generated by these operators is $`^{(\alpha )};`$ we are in the conditions of Subsections 3.1 and 3.4. We now consider the following Hilbert space: $$^{(\mu ,\alpha )}^{[\mu ]}^{(\alpha )}$$ (6.1) with $`^{[\mu ]}`$ defined by (5.11); we are denoting the mass by $`\mu >0`$ to avoid confusion with the indices $`m`$. If we identify the Fock space $`^{(\alpha )}`$ with its dual (using the scalar product $`<,>_\alpha `$) then we can describe the preceding Hilbert space as the space of Borel maps from $`^3`$ into $`^{(\alpha )}`$ square integrable with respect to the scalar product $$<f,g>=\frac{d𝐩}{2p^0}<f(p),g(p)>_\alpha .$$ (6.2) We define $$\alpha _0^jp^jj=1,\mathrm{},D2$$ (6.3) and we can define the Virasoro generators by $$L_m^{(\alpha )}\frac{1}{2}\underset{n}{}:\alpha _{mn}^j\alpha _n^j:a\delta _mI$$ (6.4) (i.e. we are in the conditions from Subsection 3.4 with $`ϵ_1=\mathrm{}=ϵ_{D2}=1).`$ More rigorously we have $`L_m^{(\alpha )}=I_1{\displaystyle \frac{1}{2}}({\displaystyle \underset{n0,m}{}}:\alpha _{mn}^j\alpha _n^j:)+p^j\alpha _m^jm0`$ $`L_0^{(\alpha )}=I_1{\displaystyle \underset{n>0}{}}:\alpha _n^j\alpha _n^j:+{\displaystyle \frac{1}{2}}\stackrel{~}{𝐩}^2I_2aI`$ (6.5) but we an use the more compact definition (6.4) without risk of confusion. By a slight abuse of notation we will called $`𝒟_0^{[\mu ]}𝒟_0`$ the algebraic Fock space. Then we have ###### Lemma 6.1 The operators $`E^j:^{(\mu ,\alpha )}^{(\mu ,\alpha )},j=1,\mathrm{},D2`$ are well defined on the algebraic Fock by the formulas $$E^ji\underset{m>0}{}\frac{1}{m}(\alpha _m^jL_mL_m\alpha _m^j)$$ (6.6) and are formally self-adjoint. Proof: We have noticed previously that the vectors $`\alpha _m^j\mathrm{\Psi }`$ and $`L_m\mathrm{\Psi }`$ (here $`\mathrm{\Psi }𝒟_0`$) are null for sufficiently large $`m>0`$ (see Proposition 3.5.) It follows that the sums in (6.6) are in fact finite if we consider vectors of the form $`E^j\mathrm{\Psi }`$. $`\mathrm{}`$ It is convenient to introduce the Hamiltonian operator according to $$H^{(\alpha )}\underset{n>0}{}:\alpha _n^j\alpha _n^j:=\underset{n>0}{}n(a_n^j)^+a_n^j.$$ (6.7) Now we have the main result: ###### Theorem 6.2 Let us define the following operators on $`𝒟_0:`$ $`P^\pm =p^\pm I_2P^j=p^jI_2`$ $`J^{kl}=L^{kl}I_2+I_1J^{(\alpha )kl}J^{k+}=L^{k+}I_2`$ $`J^k=L^kI_2+{\displaystyle \frac{1}{p^+}}E^kJ^+=L^+I_2;`$ (6.8) here $`k,l=1,\mathrm{},D2`$ as usual and the operators $`J^{(\alpha )kl}`$ are those defined by the first formula of (3.28). Then these operators are a representation of the Poincaré Lie algebra iff $`D=26`$ and we consider only the states from the physical Hilbert space $$_{phys}^{(\mu ,\alpha )}\left\{\mathrm{\Psi }^{(\mu ,\alpha )}\right|H^{(\alpha )}\mathrm{\Psi }=\left(1+\frac{\mu ^2}{2}\right)\mathrm{\Psi }\}.$$ (6.9) Proof: We have to check the formulas (5.3). Of course, we will use the fact the the operators $`L^{\mathrm{}}`$ verify these relations as stated in the previous Section so the non-trivial ones must have at least a $`J^k`$ entry. We are left with the following non-trivial relations to check: $`[J^+,J^k]=iJ^k[J^{j+},J^k]=iJ^{jk}i\delta _{jk}J^+`$ $`[J^j,J^{kl}]=i\delta _{jk}J^l+i\delta _{jl}J^k[J^j,J^k]=0.`$ (6.10) The first three preceding relations can be checked elementary and do not produce anomalies. Only the last relation is highly non-trivial. To compute the commutator $`[J^j,J^k]=0`$ we use the elementary formula $$[AB,CD]=A[B,C]D+AC[B,D]+[A,C]DB+C[A,D]B.$$ (6.11) We need to compute first $$[E^j,E^k]=(C_1^{jk}+C_2^{jk}H.c.)$$ (6.12) where $$C_1^{jk}=\underset{m,n>0}{}\frac{1}{mn}[\alpha _m^jL_m,\alpha _n^kL_n]C_2^{jk}=\underset{m,n>0}{}\frac{1}{mn}[\alpha _m^jL_m,L_n\alpha _n^k]$$ (6.13) and we understand that all operators are acting on vectors from the algebraic Fock space. These commutators can be computed using the formulas (6.11) and (3.15) with $`DD2.`$ One checks that at every stage of the computations the sums are in fact finite because we apply the commutators only on vectors from the algebraic Fock space. We give only the final results: $`C_1^{jk}=({\displaystyle \underset{m>0}{}}{\displaystyle \frac{m1}{2}}\alpha _m^j\alpha _m^k+{\displaystyle \underset{m>n0}{}}{\displaystyle \frac{1}{m}}\alpha _m^jL_{mn}\alpha _n^k)(jk)`$ $`C_2^{jk}={\displaystyle \underset{mn>0}{}}{\displaystyle \frac{1}{m}}\alpha _m^jL_{mn}\alpha _n^k+{\displaystyle \underset{nm>0}{}}{\displaystyle \frac{1}{n}}\alpha _m^jL_{mn}\alpha _n^k`$ $`+{\displaystyle \underset{m>0}{}}\left[m{\displaystyle \frac{D26}{2}}+{\displaystyle \frac{1}{m}}\left(2a{\displaystyle \frac{D2}{12}}\right)+1\right]\alpha _m^k\alpha _m^j+\mathrm{}`$ (6.14) where by $`\mathrm{}`$ we mean a term proportional to $`\delta _{jk}`$ which disappears from (6.12). We finally obtain $`[E^j,E^k]={\displaystyle \underset{m>0}{}}[m{\displaystyle \frac{D26}{12}}+{\displaystyle \frac{1}{m}}(2a{\displaystyle \frac{D2}{12}})][\alpha _m^j\alpha _m^k(jk)]`$ $`+i(p^jE^kp^kE^j)+2{\displaystyle \underset{m>0}{}}{\displaystyle \frac{1}{m}}[\alpha _m^j\alpha _m^k(jk)]L_0.`$ (6.15) Let us note that if we use the definition (6.5) we have: $`L_0=H^{(\alpha )}+\frac{1}{2}\stackrel{~}{𝐩}^2a`$ and we observe that the constant $`a`$ drops out! Now we insert the commutator (6.15) in the formula $`[J^j,J^k]=[L^j,L^k]I_2+{\displaystyle \frac{1}{(p^+)^2}}[E^j,E^k]`$ $`+\left(L^j{\displaystyle \frac{1}{p^+}}\right)E^k\left(L^k{\displaystyle \frac{1}{p^+}}\right)E^j+{\displaystyle \frac{1}{p^+}}(L^jE^k){\displaystyle \frac{1}{p^+}}(L^kE^j)`$ (6.16) and get the final result $$[J^j,J^k]=\frac{1}{(p^+)^2}\underset{m>0}{}[\alpha _m^k\alpha _m^j(jk)][m\frac{D26}{12}+\frac{1}{m}(2H^{(\alpha )}\frac{D2}{12}\mu ^2)];$$ (6.17) equating to zero we obtain the value of $`D`$ and also the expression for the physical Hilbert space. It remains to show that the physical Hilbert space is Poincaré invariant i.e. it is left invariant by the operators (6.8) from the statement. This follows from $$[H^{(\alpha )},K^j]=[L_0,K^j]=0$$ (6.18) and the proof is finished. $`\mathrm{}`$ Let us remark that the vacuum $`\mathrm{\Omega }`$ does not belong to the physical Hilbert space. The preceding system seems to be closest to what the physical intuition tells us a vibrating string of mass $`\mu `$ should be: the first factor in (6.1) describes the translation of the string in space-time and the second factor the vibrations of the string in the rest frame. Because the operator $`H^{(\alpha )}`$ has the spectrum $`\sigma (H^{(\alpha )})=`$ we obtain that in this case the mass $`\mu `$ is quantized: $`\mu ^22.`$ However, a different construction is preferred in the literature and is called the Bosonic string. Instead of (6.1) we take $$^{(b)}\left(_{lL}^{[\mu _l]}\right)^{(\alpha )};$$ (6.19) where we the sum is over an unspecified set $`L`$ (not necessarily finite) of masses. The extensions of the formulas (6.8) to this case are obvious. (In fact, $`^{(b)}`$ is a direct sum of Hilbert spaces of the type $`^{(\mu ,\alpha )}).`$ The same computation as above leads to $$[J^j,J^k]=\frac{1}{(p^+)^2}\underset{m>0}{}[\alpha _m^k\alpha _m^j(jk)][m\frac{D26}{12}+\frac{1}{m}(2H^{(\alpha )}\frac{D2}{12}p^2)]$$ (6.20) where now $$p^2=_{lL}\mu _l^2I_l.$$ (6.21) We obtain as before $`D=26`$ but the physical Hilbert space is $$_{phys}^{(b)}\left\{\mathrm{\Psi }^{(b)}\right|2(H^{(\alpha )}1)\mathrm{\Psi }=p^2\mathrm{\Psi }\}.$$ (6.22) In this way we get tachyons in the spectrum of the model (for instance the vacuum state corresponds to $`p^2=2`$). In this case the expression of $`J^{kl}`$ from (6.8) makes sense only on functions defined in the chart $`V_{D1}`$ and of compact support (such that the singularity in $`p^+=0`$ is integrable). Similar constructions must be considered in all charts. We note in the end that the necessity of considering only states lying in the physical Hilbert space (6.9) or (6.22) appears in the standard literature in a different form e.g. equation (2.3.12) from . We also note that the condition $`a=1`$ appearing frequently in the literature is not needed. Apparently this fact is known in the literature but we cannot provide an explicit reference on this point. ### 6.2 The Neveau-Schwartz Superstring We generalize the previous arguments for the superstring. In the NS case we consider the Hilbert space generated by the system of Bose oscillators $`\alpha _m^j,m^{},j=1,\mathrm{},D2`$ (i.e. we exclude for the moment the value $`m=0`$) and the the Fermi oscillators $`b_r^j,r\frac{1}{2}+,j=1,\mathrm{},D2;`$ The Hilbert space generated by these operators is $`^{(NS)};`$ we are in the conditions of Subsections 3.2 and 3.4. We now consider the following Hilbert space: $$^{(NS)}^{[\mu ]}^{(NS)}$$ (6.23) with $`^{[\mu ]}`$ defined by (5.11). We define as in the previous Subsection $$\alpha _0^jp^jj=1,\mathrm{},D2$$ (6.24) and we are in the conditions of Subsection 3.4 so we can define the Virasoro generators $$L_m^{(\alpha )}\frac{1}{2}\underset{n}{}:\alpha _{mn}^j\alpha _n^j:+\frac{1}{2}\underset{r1/2+}{}r:b_r^jb_{m+r}^j:a\delta _mI.$$ (6.25) Then we can define the operators $`E^j,F^j:^{(NS)}^{(NS)},j=1,\mathrm{},D2`$ on the algebraic Fock space by the formulas $$E^ji\underset{m>0}{}\frac{1}{m}(\alpha _m^jL_mL_m\alpha _m^j)$$ (6.26) $$F^ji\underset{m}{}K_m^{jl}\alpha _m^l=i\underset{m>0}{}(\alpha _m^lK_m^{jl}+K_m^{jl}\alpha _m^l)iK_0^{jl}\alpha _0^l$$ (6.27) where the operators $`K_m^{jl}`$ have been defined in the Subsection 4. We remark that the expression (6.26) formally coincides with (6.6) but the expression of $`L_m`$ is in fact different: we have a Fermi contribution in (6.25). $$K^jE^j+F^j$$ (6.28) and all operators $`E^j,F^j,K^j)`$ are formally self-adjoint. The Hamiltonian operator also has a Fermi contribution: $$H^{(NS)}\underset{n>0}{}:\alpha _n^j\alpha _n^j:+\underset{r>0}{}r:b_r^jb_r^j:=\underset{n>0}{}n(a_n^j)^+a_n^j+\underset{r>0}{}r:b_r^jb_r^j:$$ (6.29) Now we have the main result: ###### Theorem 6.3 Let us define the following operators on $`𝒟^{[m]}𝒟_0:`$ $`P^\pm =p^\pm I_2P^j=p^jI_2`$ $`J^{kl}=L^{kl}I_2+I_1J^{(NS)kl}J^{k+}=L^{k+}I_2`$ $`J^k=L^kI_2+{\displaystyle \frac{1}{p^+}}K^kJ^+=L^+I_2;`$ (6.30) here $`k,l=1,\mathrm{},D2`$ as usual and the operators $`J^{(NS)kl}=J^{(\alpha )kl}+J^{(b)kl}`$ are those defined by the formula (3.28). Then these operators are a representation of the Poincaré Lie algebra iff $`D=10`$ and we consider only the states from the physical Hilbert space $$_{phys}^{(NS)}\left\{\mathrm{\Psi }^{(NS)}\right|H^{(NS)}\mathrm{\Psi }=\frac{1}{2}(1+\mu ^2)\mathrm{\Psi }\}.$$ (6.31) Proof: As in the previous Subsection we check the formulas (5.3) and the obstructions can come only from the commutator $`[J^j,J^k].`$ The commutator $`[E^j,E^k]`$ can be obtained from the corresponding formula of the preceding Subsection with the substitution $`\frac{D2}{12}\frac{D2}{8}`$ in the commutators of $`L_m`$’s as it follows by comparing (3.15) to (3.27). We get in this way easily: $`[E^j,E^k]={\displaystyle \underset{m>0}{}}[m{\displaystyle \frac{D10}{8}}+{\displaystyle \frac{1}{m}}(2a{\displaystyle \frac{D2}{8}})]\alpha _m^k\alpha _m^j(jk)`$ $`+i(p^jE^kp^kE^j)+2{\displaystyle \underset{m>0}{}}{\displaystyle \frac{1}{m}}[\alpha _m^j\alpha _m^k(jk)]L_0.`$ (6.32) To obtain the expression $`[K^j,K^k]`$ we still have to compute the commutators $`[F^j,F^k]`$ and $`[E^j,F^k]`$ for which we use again (6.11). After a tedious but straightforward algebra we get $`[J^j,J^k]={\displaystyle \frac{1}{(p^+)^2}}{\displaystyle \underset{m>0}{}}[\alpha _m^k\alpha _m^j(jk)][m{\displaystyle \frac{D10}{8}}+{\displaystyle \frac{1}{m}}(2H^{(NS)}{\displaystyle \frac{D2}{8}}\mu ^2)]`$ $`+{\displaystyle \frac{D10}{4}}{\displaystyle \underset{r>0}{}}r(2r1)[b_r^jb_r^k(jk)]+{\displaystyle \frac{1}{p^+}}(2H^{(NS)}1\mu ^2)K_0^{jk};`$ (6.33) to obtain this formula we use in an essential way the formula (4.15) of Section 4. Now we equate to zero the right hand side and the theorem follows. $`\mathrm{}`$ The vacuum $`\mathrm{\Omega }`$ does not belong to the physical Hilbert space in this case also. As in the preceding Subsection, a different construction is preferred in the literature and is called the Neveau-Schwartz superstring. Instead of (6.23) we take $$^{(NS)}\left(_{lL}^{[\mu _l]}\right)^{(NS)};$$ (6.34) where we the sum is over an unspecified set $`L`$ of masses. The extensions of the formulas (6.30) to this case are obvious. The same computation as above leads to $`[J^j,J^k]={\displaystyle \frac{1}{(p^+)^2}}{\displaystyle \underset{m>0}{}}[\alpha _m^k\alpha _m^j(jk)][m{\displaystyle \frac{D10}{8}}+{\displaystyle \frac{1}{m}}(2H^{(NS)}{\displaystyle \frac{D2}{8}}p^2)]`$ $`+{\displaystyle \frac{D10}{4}}{\displaystyle \underset{r>0}{}}r(2r1)[b_r^jb_r^k(jk)]+{\displaystyle \frac{1}{p^+}}(2H^{(NS)}1p^2)K_0^{jk};`$ (6.35) where now $$p^2=_{lL}\mu _l^2I_l.$$ (6.36) We obtain as before $`D=10`$ but the physical Hilbert space is $$_{phys}^{(NS)}\left\{\mathrm{\Psi }^{(NS)}\right|(2H^{(NS)}1)\mathrm{\Psi }=p^2\mathrm{\Psi }\}.$$ (6.37) In this way we get tachyons in the spectrum of the model (for instance the vacuum state corresponds to $`p^2=1`$). One can eliminate the tachyons imposing the GSO condition namely, considering that the physical Hilbert space is the subspace of $`_{phys}^{(NS)}`$ generated by odd numbers of $`b`$ oscillators and arbitrary numbers of $`\alpha `$ oscillators; this subspace is again Poincaré invariant. The parameter $`a`$ remains unconstrained in this case also. ### 6.3 The Ramond Superstring In the Ramond case we consider the Hilbert space generated by the system of Bose oscillators $`\alpha _m^j,m^{},j=1,\mathrm{},D2`$ (i.e. we exclude for the moment the value $`m=0`$) and the the Fermi oscillators $`d_m^j,m,j=1,\mathrm{},D2;`$ the Hilbert space generated by these operators is $`^{(R)};`$ we are in the conditions of Subsections 3.2 and 3.4. We consider the following Hilbert space: $$^{(R)}^{[\mu ]}^{(R)}$$ (6.38) with $`^{[\mu ]}`$ defined by (5.11). We define $$\alpha _0^jp^jj=1,\mathrm{},D2$$ (6.39) and we are in the conditions of Subsection 3.4 so we can define the Virasoro generators $$L_m^{(\alpha )}\frac{1}{2}\underset{n}{}:\alpha _{mn}^j\alpha _n^j:+\frac{1}{2}\underset{n}{}n:d_n^jd_{m+n}^j:+\left(\frac{D}{16}a\right)\delta _mI;$$ (6.40) we remark that we have included the shift (3.23) of $`L_0`$ such that we have the canonical form for the $`2`$-cocycle of the Virasoro algebra. Then we can define the operators $`E^j,F^j,K^j:^{(NS)}^{(NS)},j=1,\mathrm{},D2`$ on the algebraic Fock by the same formulas as in the preceding Subsection (however the Virasoro operators are different). In this case the Hamiltonian operator also has a Fermi contribution: $$H^{(R)}\underset{n>0}{}:\alpha _n^j\alpha _n^j:+\underset{n>0}{}n:d_n^jd_n^j:=\underset{n>0}{}n(a_n^j)^+a_n^j+\underset{n>0}{}n:d_n^jd_n^j:$$ (6.41) Now the main result is: ###### Theorem 6.4 Let us define the following operators on $`𝒟^{[m]}𝒟_0:`$ $`P^\pm =p^\pm I_2P^j=p^jI_2`$ $`J^{kl}=L^{kl}I_2+I_1J^{(R)kl}J^{k+}=L^{k+}I_2`$ $`J^k=L^kI_2+{\displaystyle \frac{1}{p^+}}K^kJ^+=L^+I_2;`$ (6.42) here $`k,l=1,\mathrm{},D2`$ as usual and the operators $`J^{(R)kl}=J^{(\alpha )kl}+J^{(d)kl}`$ are those defined by the formula (3.28). Then these operators are a representation of the Poincaré Lie algebra iff $`D=10`$ and we consider only the states from the physical Hilbert space $$_{phys}^{(R)}\left\{\mathrm{\Psi }^{(NS)}\right|H^{(NS)}\mathrm{\Psi }=\frac{\mu ^2}{2}\mathrm{\Psi }\}.$$ (6.43) Proof: Formally, the content of this theorem coincides with the previous theorem. Similar computations, making use of (4.17) this time, lead to: $`[J^j,J^k]={\displaystyle \frac{1}{(p^+)^2}}{\displaystyle \underset{m>0}{}}[\alpha _m^k\alpha _m^j(jk)][m{\displaystyle \frac{D10}{8}}+{\displaystyle \frac{1}{m}}(2H^{(R)}\mu ^2)]`$ $`+{\displaystyle \frac{D10}{2}}{\displaystyle \underset{n>0}{}}n^2[d_n^jd_n^k(jk)]+{\displaystyle \frac{1}{p^+}}(2H^{(R)}\mu ^2)K_0^{jk}`$ (6.44) and equating to zero the right hand side and the theorem follows. $`\mathrm{}`$ In the Ramond model the vacuum $`\mathrm{\Omega }`$ belongs to the physical Hilbert space. As in the preceding Subsection, a different construction is preferred in the literature and is called the Ramond superstring. Instead of (6.38) we take $$^{(R)}\left(_{lL}^{[\mu _l]}\right)^{(R)};$$ (6.45) where we the sum is over an unspecified set $`L`$ of masses. The extensions of the formulas (6.42) to this case are obvious. We obtain as before $`D=10`$ but the physical Hilbert space is $$_{phys}^{(R)}\left\{\mathrm{\Psi }^{(R)}\right|H^{(R)}\mathrm{\Psi }=\frac{p^2}{2}\mathrm{\Psi }\}.$$ (6.46) In this way we do not get tachyons in the spectrum of the model. ### 6.4 Other Superstring Models From the preceding two Subsections it is clear that formulas of the type (4.15) and (4.17) are essential for establishing Lorentz invariance. We investigate now if such formulas can be valid for more general cases. More precisely, suppose that we have a $`N`$-dimensional representation $`\sigma _{jk},j,k=1,\mathrm{},D`$ of the algebra $`so(D)`$ such that $$\sigma _{jk}^T=\sigma _{jk};$$ (6.47) then we can define the associated Kac-Moody algebras $`K_m^{jk}(\sigma )`$ according to the formulas (4.5) and (4.16) respectively. We are interested if in some special cases formulas of the type (4.15) and (4.17) hold. A necessary condition is that the terms quadratic in the operators $`b`$ (resp. $`d`$) cancel identically. It is not very difficult to prove that this requirement is equivalent to $$\underset{r_1+\mathrm{}+r_4=0}{}X_{p_1r_1;\mathrm{};p_4r_4}^{jk}:b_{r_1}^{p_1}\mathrm{}b_{r_4}^{p_4}:=0$$ (6.48) where $$X_{p_1r_1;\mathrm{};p_4r_4}^{jk}(r_1+r_2)(\sigma _{jl})_{p_1p_2}(\sigma _{kl})_{p_3p_4}+(r_3r_4)(\sigma _{jk})_{p_1p_2}\delta _{p_3p_4}$$ (6.49) in the Neveau-Schwartz case and similar relations for the Ramond case. The relation (6.48) is equivalent to $$X_{p_1r_1;\mathrm{};p_4r_4}^{jk}(13)(14)(23)(24)+(13,24)=0.$$ (6.50) One inserts here the definition (6.49) and eliminates $`r_4=(r_1+r_2+r_3);`$ the result is an equation of the form $$r_1E_{p_1\mathrm{};p_4}^{(1)jk}+r_2E_{p_1\mathrm{};p_4}^{(2)jk}+r_3E_{p_1\mathrm{};p_4}^{(3)jk}=0$$ (6.51) so we obtain three relations $$E_{p_1\mathrm{};p_4}^{(a)jk}=0a=1,2,3.$$ (6.52) One can easily see that the relation $$E^{(1)}+E^{(2)}E^{(3)}=0$$ (6.53) is equivalent to $$(\sigma _{jl})_{ab}(\sigma _{kl})_{cd}(jk)=\delta _{bd}(\sigma _{jk})_{ac}(ab)(cd)+(ab,cd)$$ (6.54) and conversely (6.54) is implies (6.52). Moreover, if we have (6.54) the one can prove that relations of the type (4.15) and (4.17) hold and we have Lorentz invariance theorems like in the preceding Subsection in $`10`$ dimensions. So the key relation (6.54) must be analyzed in the case $`D=8`$. We note that in this case one should modify in an appropriate way the expression (3.28) for the generators of the rotations $`J^{(b)jk}`$ and $`J^{(d)jk}`$. One can obtain an important consequence of (6.54) if we take $`b=c`$ and sum over $`b=1,\mathrm{},N`$. We obtain $$[\sigma _{jl},\sigma _{kl}]=(2N)\sigma _{jk};$$ (6.55) on the other hand we have from the representation property of $`\sigma _{jk}`$ $$[\sigma _{jl},\sigma _{kl}]=(2D)\sigma _{jk}$$ (6.56) so we conclude that the representation $`\sigma _{jk}`$ should be $`D`$-dimensional, i.e. we need to consider only the representations of dimension $`8`$ of the algebra $`so(8)`$. It is known that there are four (non-equivalent) such representations: the vector representation (which we have already used in the preceding Subsection), the adjoint representation and two the spinor representations $`\mathrm{𝟖}_s`$ and $`\mathrm{𝟖}_c`$ of opposite chirality. It seems that the identity (6.54) is valid for the for the spinor representations also but the details are not easily found in the literature so we provide an elementary analysis. First, it is clear that if we multiply (6.54) with $`M_{dc}`$ and make the summation we obtain an equivalent relation $$Tr(\sigma _{jl}M)\sigma _{kl}(jk)=[MM^T,\sigma _{jk}],M.$$ (6.57) If $`M`$ is symmetric then the preceding relation is an identity. So (6.54) is equivalent to $$Tr(\sigma _{jl}M)\sigma _{kl}(jk)=2[M,\sigma _{jk}]$$ (6.58) for all antisymmetric $`N\times N`$\- matrices $`M`$. Now the number of $`N\times N`$\- antisymmetric matrices is $`\frac{N(N1)}{2};`$ on the other hand the number of matrices $`\sigma _{jk}`$ is $`\frac{D(D1)}{2}.`$ But we have already established that $`N=D`$ so if the matrices $`\sigma _{jk}`$ are linear independent the relation (6.58) is equivalent to $$Tr(\sigma _{jl}\sigma _{pq})\sigma _{kl}(jk)=2[\sigma _{pq},\sigma _{jk}],p,q.$$ (6.59) In particular is is easy to see that the fundamental representation $`E_{jk}`$ verifies the preceding identity. We check the identity for the spinor representations. According to one can describe the spinor representations of the algebra $`so(2n)`$ considering the Fermi Fock space $`S`$ generated by the operators $`b_j,b_j^{},j=1,\mathrm{},n;`$ we have the CAR algebra: $$\{b_j,b_k\}=0\{b_j,b_k^{}\}=\delta _{jk}.$$ (6.60) Next, we define the operators $`\gamma _j=b_j+b_j^{},j=1,\mathrm{},n`$ $`\gamma _j=i(b_{nj}b_{nj}^{}),j=n+1,\mathrm{},2n`$ (6.61) and prove immediately that they form a $`2^n`$\- dimensional representation of the Clifford algebra $`C(2n,0)`$ i.e. we have $$\{\gamma _j,\gamma _k\}=\delta _{jk}I.$$ (6.62) Then a representation of the algebra $`so(2n)`$ is given by the operators $$\sigma _{jk}=\frac{1}{4}[\gamma _j,\gamma _k].$$ (6.63) This representation is not irreducible. In fact let us denote by $`S_+`$ (resp. $`S_{})`$ the subspaces of $`S`$ generated by applying an even (resp. odd) number of creation operators $`b_j^{}`$ on the vacuum. The projectors on these subspaces will be denoted by $`P_\pm `$. It is easy to see that these two subspaces are left invariant by the representation $`\sigma _{jk}`$ so it makes sense to define the restrictions $`\sigma _{jk}^\pm .`$ The operators $`\sigma _{jk}^\pm `$ are immediately seen to be linear independent. It is also easy to prove that $`dim(S_+)=dim(S_{})=2^{n1}`$ i.e. both representations $`\sigma _{jk}^\pm .`$ are of the same dimension $`2^{n1}`$; these are, by definition, the spinor representations of the algebra $`so(2n)`$. Finally we prove: $$Tr(\sigma _{jk}^\pm \sigma _{pq}^\pm )=2^{n3}(\delta _{kp}\delta _{jq}\delta _{jp}\delta _{kq}).$$ (6.64) Indeed, because the left hand side is a $`SO(2n)`$\- invariant tensor and because of the antisymmetry properties we know that the right hand side must have the form $`\lambda (\delta _{kp}\delta _{jq}\delta _{jp}\delta _{kq});`$ to determine the constant $`\lambda `$ we consider a particular case, say $`p=j,q=kjk`$ and we obtain $$\lambda =Tr(\sigma _{jk}^2P_\pm )=\frac{1}{4}Tr(P_\pm )=\frac{1}{4}dim(S_\pm )=2^{n3}.$$ (6.65) It follows that only for $`n=4`$ we have $`\lambda =2`$ and in this case if we use (6.64) in (6.59) we obtain an identity. It follows that the two spinor representations of $`so(8)`$ verify the identity (6.58) so they can be used to construct supersymmetric string models as in the preceding two Subsections. These models are considered more consistent because we quantize using Fermi statistics oscillators pertaining to spinor representations so we are in agreement with spin-statistics correspondence. Moreover these models exhibit also supersymmetry invariance. There is yet another possibility of constructing consistent models, namely by modifying the Bosonic string from Subsection 6.1. We consider that we have another representation of the Virasoro algebra $`L_m^c`$ with central charge $`c`$ acting in the Hilbert space $`^c;`$ we consider the Hilbert space $`^{\mu ,\alpha }^c`$ where $`^{\mu ,\alpha }`$ is given by (6.1) and modify the Virasoro algebra given by (6.4) according to $`L_mL_m+L_m^c.`$ Because the central charges are additive the new Virasoro algebra will have the central charge $`D2+c`$ so the consistency condition is in this case $$c=26D$$ (6.66) and the expression of the physical Hilbert space from theorem 6.9 remains the same. In particular if we want a model in $`D=10`$ dimensions we must find out a representation of the Virasoro algebra with central charge $`c=16`$. It is known that such representations can be found for the groups $`SO(32)`$ and $`E_8\times E_8`$ using Sugawara construction . This new possibility is used in the construction of the heterotic string models. For the description of closed strings a doubling of the Bose oscillators $`\alpha _m^j`$ appears corresponding to the left and right oscillator modes. Composing in various ways the models described above one can obtain the well-known string models of type I, IIA, IIB and heterotic. ## 7 Covariant Quantization of Strings and Superstrings One can construct a manifestly covariant formalism also . The idea is to take in Subsection 3.4 the case $`ϵ_0=1,ϵ_1=\mathrm{}=ϵ_{D1}=1.`$ In this way the Hilbert space will have states of negative or zero norm. So we consider that we have the family of operators: $`\alpha _m^\mu ,m^{},\mu =0,\mathrm{},D`$ acting in the Hilbert space $`_{\mathrm{cov}}^{(\alpha )}`$ such that: $`[\alpha _m^\mu ,\alpha _n^\nu ]=\eta _{\mu \nu }m\delta _{m+n}I,m,n`$ $`\alpha _m^\mu \mathrm{\Omega }=0,m>0`$ $`(\alpha _m^\mu )^+=\alpha _m^\mu m;`$ (7.1) this Hilbert space will not be positively defined. Define the Virasoro operators $$\overline{L}_m\frac{1}{2}\eta _{\mu \nu }\underset{n0,m}{}:\alpha _{mn}^\mu \alpha _n^\nu :$$ (7.2) and we have the following commutation relations: $$[\overline{L}_m,\overline{L}_n]=(mn)\overline{L}^{m+n}+D\frac{m(m^21)}{12}\delta _{m+n}I.$$ (7.3) ###### Proposition 7.1 We consider $`k^D`$ and recursively define the operators $`U_n(k),n`$ according to $$U_0=IU_n(k)=\frac{1}{n}\underset{l=1}{\overset{n}{}}U_{nl}(k)k\alpha _l.$$ (7.4) For convenience we define $`U_n=0n<0.`$ Then the following relation is valid: $$[\alpha _m^\mu ,U_n(k)]=\theta (m1)k^\mu U_{m+n}(k)$$ (7.5) where $`\theta (m)`$ is the usual Heaviside function. The proof is easily done by induction on $`n`$. Let us remark that the expressions $`U_n(k)`$ are the coefficients of the formal series in $`z:`$ $$U(z,k)e^{A(z,k)}A(z,k)\underset{n1}{}\frac{1}{n}k\alpha _nz^n.$$ (7.6) The recurrence relation from the statement of the proposition can be found if we compute the derivative of $$U(z,k)=\underset{n1}{}U_n(k)z^n$$ (7.7) in two ways. The explicit relation is $$U_n(k)=\underset{p0}{}\frac{1}{p!}\underset{i_1,\mathrm{},i_p>0}{}\underset{i_1+\mathrm{}i_p=n}{}\frac{1}{i_1\mathrm{}i_p}(k\alpha _{i_1})\mathrm{}(k\alpha _{i_p})$$ (7.8) but it is convenient to work with the recurrence relation and not with the explicit expression given above. The operators $`U_n(k)`$ are leaving the algebraic Fock space $`𝒟_0`$ invariant and moreover for every $`\mathrm{\Psi }𝒟_0`$ we have $$U_n(k)\mathrm{\Psi }=0$$ (7.9) for sufficiently large $`n`$. It is useful to expressed the formal series relations: $`[U(z,k)^{},U(z,k^{})]=0\mathrm{for}kk^{}=0`$ $`U(z,k)U(z,k^{})=U(z,k+k^{}).`$ (7.10) in terms of the $`U_p(k)`$ operators. In particular we have $$\underset{p}{}U_{np}(k)U_p(k^{})=U_n(k+k^{}).$$ (7.11) Less elementary are the following sum relations: $`{\displaystyle \underset{p}{}}pU_{np}(k)U_{m+p}(k^{})={\displaystyle \underset{l>0}{}}U_{m+nl}(k+k^{})k^{}\alpha _lmU_{m+n}(k+k^{})`$ $`{\displaystyle \underset{p}{}}pU_{np}(k)U_{m+p}(k)=\theta (m+n)k\alpha _{m+n}m\delta _{m+n}I`$ $`{\displaystyle \underset{p}{}}pU_{np}(\alpha k)U_{m+p}(\beta k^{})={\displaystyle \frac{n\beta m\alpha }{\alpha +\beta }}U_{m+n}((\alpha +\beta )k)\alpha ,\beta ,\alpha +\beta 0.`$ (7.12) Now we have ###### Proposition 7.2 The operators $`V_n(k),n`$ are well defined on the algebraic Fock space according to the relations $$V_n(k)=\underset{p}{}U_{pn}(k)^{}U_p(k).$$ (7.13) Indeed the sum over $`p`$ is in fact finite because we have $`U_n(k)\mathrm{\Psi }=0`$ for sufficiently high $`n`$ if $`\mathrm{\Psi }𝒟_0.`$ The expressions $`V_n(k)`$ are the coefficients of the formal series $$V(z,k)U(z,k)^{}U(z,k)$$ (7.14) We have analogue elementary properties: the operators $`V_n(k)`$ are leaving the algebraic Fock space $`𝒟_0`$ invariant and moreover for every $`\mathrm{\Psi }𝒟_0`$ we have $$V_n(k)\mathrm{\Psi }=0$$ (7.15) for sufficiently large $`n`$. We have $`V_n(k)^{}=V_n(k).`$ $`[\alpha _m^\mu ,V_n(k)]=(1\delta _m)k^\mu V_{m+n}(k).`$ $`{\displaystyle \underset{p}{}}V_{np}(k)V_p(k^{})=V_n(k+k^{}).`$ (7.16) Sum relations of the type (7.12) can be found for the $`V_p(k)`$ operators. Now we can derive the conformal properties of these operators i.e. the commutation relations with the Virasoro operators. ###### Proposition 7.3 The following relation is true: $$[\overline{L}_m,V_n(k)]=(c_mk^2+m+n)V_{m+n}$$ (7.17) where we have defined $$c_m\{\begin{array}{cc}\frac{m1}{2},\hfill & \text{for m }>\text{ 0}\hfill \\ \frac{m+1}{2},\hfill & \text{for m }<\text{ 0}\hfill \\ 0,\hfill & \text{for m = 0}\hfill \end{array}.$$ (7.18) In particular if $`k^2=0`$ we have $$[\overline{L}_m,V_n(k)]=(m+n)V_{m+n}$$ (7.19) i.e. the operators $`V_n(k)`$ have conformal dimension $`0`$. The computation is straightforward: we first compute the commutations relation of the Virasoro operators with $`U_n(k)`$ and then we use the definition of the operators $`V_n(k).`$ We only note a discrepancy with the standard literature where it is asserted that these operators have defined conformal dimension for any $`k^D;`$ the origin of this discrepancy can be traced to the coefficient $`c_m`$ which is different from the standard literature. Fortunately, only the case $`k^2=0`$ is needed for the construction of the DDF operators. We are approaching the definition of the DDF operators. First we define the operators $$\overline{V}_n^\mu (k)\underset{p>0}{}[\alpha _p^\mu V_{n+p}(k)+V_{np}(k)\alpha _p^\mu ]$$ (7.20) and we have ###### Proposition 7.4 Let $`k^D,k^2=0.`$ Let us define the following operators: $$\overline{A}_m^\mu \overline{V}_m^\mu (mk).$$ (7.21) Then the following relations are verified: $`[\overline{A}_m^\mu ,\overline{A}_n^\nu ]=\eta ^{\mu \nu }m\delta _{m+n}I+k^\mu \overline{V}_{m,n}^\nu (k)k^\nu \overline{V}_{n,m}^\mu (k)`$ $`[\overline{L}_m,\overline{A}_n^\mu ]=n\overline{V}_{m+n}^\mu +{\displaystyle \frac{m(m1)}{2}}k^\mu V_{m+n}(nk)`$ $`\overline{A}_n^\mu (k)^{}=\overline{A}_n^\mu (k)`$ $`\overline{A}_n^\mu \mathrm{\Omega }=0m>0`$ $`\overline{A}_0^\mu =0`$ (7.22) where the explicit expressions $`\overline{V}_{m,n}^\nu (k)`$ are not important. To construct the DDF operators we have to include the kinematic degrees of freedom also. We define the Hilbert space $`_{\mathrm{cov}}^{[\mu ,\alpha ]}^{[\mu ]}_{\mathrm{cov}}^{(\alpha )}`$ where $`^{[\mu ]}`$ is the Hilbert space of a particle of mass $`\mu `$ and spin $`0`$ and $`_{\mathrm{cov}}^{(\alpha )}`$ is the Fock space defined at the beginning of this Section. We use the convention $$\alpha _0^\mu =p^\mu $$ (7.23) and define the covariant Virasoro operators $$L_m^{(\alpha )}\frac{1}{2}\eta _{\mu \nu }\underset{n}{}:\alpha _{mn}^\mu \alpha _n^\nu :a\delta _mI$$ (7.24) such that we have the following commutation relations: $$[L_m^{(\alpha )},L_n^{(\alpha )}]=(mn)L_{m+n}^{(\alpha )}+D\frac{m(m^21)}{12}\delta _{m+n}I.$$ (7.25) In this Hilbert space we have a natural action of the Poincaré algebra. This Hilbert space will have states of negative or zero norm. Now we can define the DDF operators: ###### Theorem 7.5 Let $`k^D`$ be such that $`k^2=0.`$ Let us define in $`_{\mathrm{cov}}^{[\mu ,\alpha ]}`$ the operators $$V_n^\mu (k)\overline{V}_n^\mu (k)+p^\mu V_n(k)$$ (7.26) and $$A_n^\mu V_n^\mu (nk)$$ (7.27) Then the following relations are true: $`[A_m^\mu ,A_n^\nu ]=\eta ^{\mu \nu }m\delta _{m+n}I+k^\mu V_{m,n}^\nu (k)k^\nu V_{n,m}^\mu (k)`$ $`[L_m^{(\alpha )},A_n^\mu ]=n(1+kp)V_{m+n}^\mu +{\displaystyle \frac{m(m1)}{2}}nk^\mu V_{m+n}(nk)`$ (7.28) where the explicit expressions $`V_{m,n}^\nu (k)`$ are not important. In particular consider that $`k^D`$ also verifies $`k^j=0,j=1,\mathrm{},D1,kp=1`$ (e.g. $`k^+=0,k^{}=\frac{1}{p^+},k^j=0,j=1,\mathrm{},D1`$) then the operators $`A_n^j,j=1,\mathrm{},D1`$ verify $`[A_m^j,A_n^k]=\delta _{jk}m\delta _{m+n}I`$ $`[L_m^{(\alpha )},A_n^j]=0,m0[L_0^{(\alpha )},A_n^j]=nA_n^j`$ $`(A_n^j)^{}=A_n^j`$ $`A_n^j\mathrm{\Omega }=0m>0`$ $`A_0^j=p^j`$ (7.29) so they verify the same algebra as the operators $`\alpha _m^j.`$ The DDF operators $`A_n^j`$ are the $`z`$-independent component of the vertex operator $$\dot{X}^j(z)e^{ikX(z,nk)}$$ (7.30) where $$X^\mu (z)\underset{n0}{}\frac{1}{n}\alpha _n^\mu z^n+p^\mu ln(z)\dot{X}_n^j\underset{n0}{}\alpha _n^jz^n+p^j.$$ (7.31) ## 8 The Covariant Quantum Bosonic String We describe the Bosonic string (see Section 6.1) using the Hilbert space bundle formalism . First we consider the system of Bose oscillators $`\alpha _m^j,m^{},j=1,\mathrm{},D2`$ (i.e. we exclude for the moment the value $`m=0`$). The Hilbert space generated by these operators is $`^{(\alpha )};`$ we are in the conditions of Subsections 3.4 and 6.1. We now consider the following Hilbert space: $$^{(\mu ,\alpha )}^{[\mu ]}^{(\alpha )}$$ (8.1) with $`^{[\mu ]}`$ defined as above. We define $$\alpha _0^jp^jj=1,\mathrm{},D2$$ (8.2) and the transversal Virasoro generators by $$L_m^T\frac{1}{2}\underset{n}{}:\alpha _{mn}^j\alpha _n^j:a\delta _mI$$ (8.3) which verify the Virasoro algebra with central charge $`D2.`$ Then we define similarly to (6.6) the operators $`E^j(p):^{(\mu ,\alpha )}^{(\mu ,\alpha )},j=1,\mathrm{},D2`$ on the algebraic Fock according to the formulas $$E^j(p)i\underset{m>0}{}\frac{1}{m}(\alpha _m^jL_m^TL_m^T\alpha _m^j);$$ (8.4) we can now construct the generators of the Poincaré group as in (6.8) and the physical Hilbert space $`_{\mathrm{phys}}^{(\mu ,\alpha )}`$ as in (6.9); we take $`D=26`$ such that the Poincaré algebra closes. The Hilbert space bundle $`^{[\mu ,\alpha ]}`$ is made of couples $`(p,f)`$ where $`p=(p^+,\stackrel{~}{p})_+^{D2}`$ is a chart on the mass-shell and $`f_{\mathrm{phys}}^{(\mu ,\alpha )};`$ there is a natural fibration over the mass shell given by the canonical projection on the first component. On this bundle we have the following action of the Lorentz algebra: $$\xi (p,f)=(\xi p,\xi f)\xi \mathrm{Lie}()$$ (8.5) where $$j^{\mu \nu }p=L^{\mu \nu }p$$ (8.6) and $$j^{kl}=J^{(\alpha )kl}j^{k+}=0j^k=\frac{1}{p^+}E^k(p)j^+=0;$$ (8.7) here $`j,k=1,\mathrm{},D2.`$ The scalar product in the fiber over $`p`$ is simply the scalar product from $`^{(\alpha )}.`$ It is easy to verify all the axioms of a Hilbert space bundle. As it is well-known the representations of the Poincaré group are induced by representations of the stability subgroup of any point on the mass-shell. If we take the point $`p^{(0)}`$ with coordinates $`p^+=\frac{\mu }{\sqrt{2}},p^j=0(j=1,\mathrm{},D2)`$ it is easy to get from (8.6) that the stability subgroup is $`SO(D1)`$ and the infinitesimal generators are $`j^{kl}k,l=1,\mathrm{},D1.`$ Next we get from (8.7) that the representation of $`SO(D1)`$ inducing the representation of the Poincaré group from the theorem is $$j^{kl}=J^{(\alpha )kl}j^{k+}=0j^k=\frac{1}{\mu }E^j(p^{(0)})j^+=0;$$ (8.8) one can check the representation property using the definition of $`_{\mathrm{phys}}^{(\mu ,\alpha )}.`$ We give now the covariant description of the preceding construction. We define the Hilbert space bundle $`_{\mathrm{cov}}^{[\mu ,\alpha ]}`$ of couples $`(p,\mathrm{\Psi })`$ where $`p`$ is on the positive mass-shell $`p^Dp^0>0p^2=\mu ^2`$ and $`\mathrm{\Psi }_{\mathrm{cov}}^{[\mu ,\alpha ]}`$ verifies the supplementary restrictions $$L_m\mathrm{\Psi }=0m0$$ (8.9) where the Virasoro operators $`L_m=L_m^{(\alpha )}`$ are given by (7.24) for the value $`a=1.`$ The Hermitian form in the fiber over $`p`$ is the form defined on $`_{\mathrm{cov}}^{(\alpha )}.`$ We want to obtain an isomorphism to the previously obtained $`_{phys}.`$ We present briefly the usual argument with some simplifications. We first define the DDF states as linear combinations of states of the form: $$f=A_{m_1}^{j_1}\mathrm{}A_{m_l}^{j_l}\mathrm{\Omega }m_1,\mathrm{},m_l<0.$$ (8.10) It is useful to introduce the notation $$K_mk\alpha _m$$ (8.11) and we easily obtain $$[K_m,K_n]=0[K_m,L_n]=mK_{m+n};$$ (8.12) we also have for any DDF state: $$K_mf=0m>0.$$ (8.13) Next we have the following technical result for which we present a simpler proof: ###### Proposition 8.1 The vectors of the type $$\mathrm{\Psi }_{\lambda ,\mu ,f}L_1^{\lambda _1}\mathrm{}L_m^{\lambda _m}K_1^{\mu _1}\mathrm{}K_n^{\mu _n}f$$ (8.14) where $`\lambda _1,\mathrm{},\lambda _m,\mu _1,\mathrm{},\mu _n^{}`$ and $`f`$ is a DDF state, are linearly independent and generates the whole space $`_{\mathrm{cov}}^{(\alpha )}.`$ Proof: (i) We know that the Hilbert space $`_{\mathrm{cov}}^{(\alpha )}`$ is generated by the operators $`\alpha _m^\mu ,m>0,\mu =0,\mathrm{},D1`$ applied on the vacuum. It is convenient to work with the operators $$\alpha _m^\pm \frac{1}{\sqrt{2}}(\alpha _m^0\pm \alpha _m^{D1})\alpha _m^jj=1,\mathrm{},D2.$$ (8.15) If we take $`k^D`$ as in the construction of the DDF operators i.e. $`k^+=0,k^j=0(j=1,\mathrm{},D2)`$ we have $$K_m=k^{}\alpha _m^+.$$ (8.16) So if we apply on the vacuum operators of the form $`K_1,\mathrm{},K_m`$ we obtain all the states of the form $`\mathrm{\Psi }^+=P(\alpha _1^+,\mathrm{},\alpha _m^+)\mathrm{\Omega }`$ with $`P`$ a polynomial. Now we easily compute $$A_1^j\mathrm{\Omega }=\alpha _1^j\mathrm{\Omega }+p^jV_1(k)\mathrm{\Omega }.$$ (8.17) Because the second vector is of the type $`\mathrm{\Psi }^+`$ we can generate the states $`\alpha _1^j\mathrm{\Omega }`$ using the DDF operators and the $`K_m`$ operators. Afterwards, using the $`K`$ operators we can generate all the states of the form $`\alpha _1^jP(\alpha _1^+,\mathrm{},\alpha _m^+)\mathrm{\Omega }.`$ Using now $`2,3,\mathrm{}`$ DDF operators we can establish by inductions that all states of the form $`P(\alpha _1^j,\alpha _1^+,\mathrm{},\alpha _m^+)\mathrm{\Omega }`$ can be obtained using only DDF and $`K`$ operators. Next we suppose that we can create all states of the form $`P(\alpha _1^{j_1},\mathrm{},\alpha _{(n1)}^{j_{n1}},\alpha _1^+,\mathrm{},\alpha _m^+)\mathrm{\Omega }`$ using only DDF and $`K`$ operators and extend the result to $`n`$ by applying DDF operators of the form $`A_n^j`$ on such a state. Finally, we note that we have $$L_1^{\lambda _1}\mathrm{}L_m^{\lambda _m}=\mathrm{const}(\alpha _1^{})^{\lambda _1+\mathrm{}\lambda _m}+\mathrm{}$$ (8.18) where by $`\mathrm{}`$ we mean terms containing $`\alpha _1^{}`$ at a power strictly smaller that $`\lambda _1+\mathrm{}\lambda _m.`$ If we choose the preceding sum conveniently we can generate all states with $`\alpha _1^{}`$ factors. In the same way we obtain the states with $`\alpha _l^{}l>1`$ factors. It follows that the states of the form $`\mathrm{\Psi }_{\lambda ,\mu ,f}`$ are really generating the whole Hilbert space $`_{\mathrm{cov}}^{(\alpha )}.`$ (ii) We must prove that there are no linear dependencies between such vectors. If we use the well-known relation $$[\overline{L}_0,\alpha _m^\mu ]=m\alpha _m^\mu $$ (8.19) we easily obtain that the vector $$\mathrm{\Psi }=\underset{n,\rho }{}(\alpha _n^\rho )^{ϵ_{n,\rho }}\mathrm{\Omega },$$ (8.20) where the product runs over a finite set of indices, is an eigenvector of $`\overline{L}_0`$ corresponding to the eigenvalue $$\lambda =\underset{n,\rho }{}nϵ_{n,\rho }.$$ (8.21) We will denote by $`_n`$ the eigenspace of $`\overline{L}_0`$ corresponding to the eigenvalue $`n`$. One can prove that $$_m_n=\{0\}$$ (8.22) for $`mn`$ using a Vandermonde determinant. Because the subspaces $`_n,n0`$ generate the whole $`_{\mathrm{cov}}^{(\alpha )}`$ we have the direct sum decomposition $$_{\mathrm{cov}}^{(\alpha )}=_{n0}_n.$$ (8.23) Now we use the relations $$[\overline{L}_0,L_m]=mL_m[\overline{L}_0,K_m]=mK_m[\overline{L}_0,A_m^j]=mA_m^j$$ (8.24) and find out that the vector $$\mathrm{\Psi }^{}=L_m^{\lambda _m}K_n^{\mu _n}(A_p^j)^{\beta _{p,j}}\mathrm{\Omega }$$ (8.25) is an eigenvector of $`\overline{L}_0`$ corresponding to the eigenvalue $$\lambda ^{}=\underset{n}{}n(\lambda _n+\mu _n+\underset{j}{}\beta _{n,j}).$$ (8.26) Let us fix $`N.`$ Then $`_N`$ is generated by vectors of the type (8.20) with $`\lambda =N;`$ on the other hand the vectors of the type (8.25) are also generating the whole Hilbert space but only those corresponding to $`\lambda ^{}=N`$ are in $`_N.`$ It follows that $`_N`$ is generated by the vectors of the type (8.25) corresponding to $`\lambda ^{}=N.`$ Because the (finite) number of vectors of the type (8.20) corresponding to $`\lambda =N`$ is the same as the number of the vectors of the type (8.25) corresponding to $`\lambda ^{}=N`$ it means that the vectors of the type (8.25) corresponding to $`\lambda ^{}=N`$ must be linear independent. $`\mathrm{}`$ We note that in our proof we did not have to compute the complicated determinant used in the proof from . The rest of the proof is standard and can be found in . The final result is: ###### Proposition 8.2 Let $`D=26`$ and the vector $`\mathrm{\Psi }_{\mathrm{cov}}^{(\alpha )}`$ verifying $$L_m\mathrm{\Psi }=0m0;$$ (8.27) then we can uniquely write it in the form $$\mathrm{\Psi }=f+s$$ (8.28) where $`f`$ is a DDF state, $`sS`$ and we have $$L_0f=fL_0s=sL_ms=0(m>0).$$ (8.29) The end of this analysis is: ###### Theorem 8.3 The Hermitian form on the Hilbert space bundle $`_{\mathrm{cov}}^{[\mu ,\alpha ]}`$ is positively defined. If we factor out the states of null norm we obtain a representation of the Poincaré group equivalent to the representation in the bundle $`^{[\mu ,\alpha ]}.`$ Proof: (i) If we use the preceding proposition we can write any element in the fiber over $`p`$ as $`\mathrm{\Psi }=f+s`$ where $`f`$ is a DDF state, $`sS`$ and we have the relations (8.29). From these relations it easily follows that we have $$<\mathrm{\Psi },\mathrm{\Psi }>=<f,f>0$$ (8.30) so if we eliminate the null-norm states we end up with a factor Hilbert space bundle with fibres isomorphic to the subspace of DDF states. (ii) We have to determine the representation of the stability subgroup $`SO(D1)`$ of the point $`p^{(0)}.`$ It is clear that we have $$J^{(\alpha )kl}A_n^j=i(\delta _{kj}A_n^l\delta _{lj}A_n^k)j,k,l=1,\mathrm{},D2$$ (8.31) so we have for any DDF state $$J^{(\alpha )kl}f=j^{(\alpha )kl}f.$$ (8.32) We still have to compute the action of the generators $`J^{(\alpha )k,D1}`$ on the fiber. It is important to note that from the first relation (8.29) we have $$\alpha _m^{}f=\frac{\sqrt{2}}{\mu }\overline{L}_mfm>0;$$ (8.33) also it is easy to prove that $$\alpha _m^+f=0m>0.$$ (8.34) Using these relations it follows that for any two DDF states $`f,f^{}`$ we have $$<f^{},J^{(\alpha )k,D1}f>=<f^{},j^{(\alpha )k,D1}f>$$ (8.35) where in the right-hand side we have the operators (8.7). It is more complicated to extend this relation for $`f^{}\mathrm{\Psi }_{\mathrm{cov}}^{(\alpha )};`$ for this we have to use the generic form of states $`\mathrm{\Psi }_{\lambda ,\mu ,f}`$ and commute $`L_m`$ and $`K_m`$ with $`E^j(p^{(0)}).`$ As a result we have for any DDF state $$J^{(\alpha )k,D1}f=j^{(\alpha )k,D1}f.$$ (8.36) Next we note that we have for any DDF state $$\overline{L}_0f=\left(1+\frac{\mu ^2}{2}\right)f;$$ (8.37) from here it follows that $$<f^{},\overline{L}_0^Tf>=\left(1+\frac{\mu ^2}{2}\right)<f^{},f>$$ (8.38) where $`\overline{L}_0^T=L_0^T`$ is the transversal part of $`\overline{L}_0`$ (i.e. it contains only the modes $`1,\mathrm{},D2`$). As above we can extend the relation for $`f^{}\mathrm{\Psi }_{\mathrm{cov}}^{(\alpha )}`$ so we have $$L_0^Tf=\left(1+\frac{\mu ^2}{2}\right)f$$ (8.39) for any DDF state. From (8.36) and (8.39) it follows that the fiber over $`p^{(0)}`$ of the fiber bundle $`_{\mathrm{cov}}^{[\mu ,\alpha ]}`$ coincides with the fiber over the same point of the fiber bundle $`^{[\mu ,\alpha ]}.`$ According to a standard theorem 9.20 of it follows that the two representations of the Poincaré group are equivalent. $`\mathrm{}`$ ## 9 BRST Quantization of the Bosonic String Another possibility is to introduce ghost degrees of freedom and construct a gauge charge operator $`Q`$ which squares to zero $`Q^2=0`$ in such a way that there is a canonical isomorphism between the physical Hilbert space and the factor space $`Ker(Q)/Im(Q)`$ , , , , . We provide here an elementary treatment. First we define the ghost Hilbert space $`_1^{gh};`$ by definition it is generated by the operators $`b_m,c_mm`$ from the vacuum $`\mathrm{\Omega }_{gh}_1^{gh}`$; we assume that $$b_m\mathrm{\Omega }_{gh}=0c_m\mathrm{\Omega }_{gh}=0m>0.$$ (9.1) These operators are subject to the following anticommutation relations: $$\{b_m,b_n\}=0\{c_m,c_n\}=0\{b_m,c_n\}=\delta _{m+n}I;$$ (9.2) we also suppose that there is a conjugation operation in $`_1^{gh}`$ such that $$b_m^{}=b_mc_m^{}=c_m.$$ (9.3) We can give a concrete realization as follows: $`_1^{gh}=_{b,c}𝒞`$ where $`_{b,c}`$ is the Fock space generated by the operators $`b_m,c_mm^{}`$ and $`𝒞`$ is the Clifford algebra generated by $`b_0,c_0.`$ ###### Proposition 9.1 The following operators $$l_m^{(1)}=\underset{n}{}(m+n):b_{mn}c_n:$$ (9.4) are well defined on the algebraic Hilbert space and are verifying: $`[l_m^{(1)},b_n]=(mn)b_{m+n}[l_m^{(1)},c_n]=(2m+n)b_{m+n}`$ $`[l_m^{(1)},l_n^{(1)}]=(mn)l_{m+n}^{(1)}+{\displaystyle \frac{1}{6}}m(113m^2)\delta _{m+n}I`$ $`(l_m^{(1)})^{}=l_m^{(1)}.`$ (9.5) Proof: We write $$l_m^{(1)}=l_m^{}+mb_mc_0+2mb_0c_m$$ (9.6) where $`l_m^{}`$ contains only the non-zero modes: $$l_m^{}=\underset{n0,m}{}(m+n):b_{mn}c_n:$$ (9.7) For the non-zero modes the $`2`$-point functions are $`<\mathrm{\Omega }_{gh},b_mc_n\mathrm{\Omega }_{gh}>=\theta (m)\delta _{m+n}<\mathrm{\Omega }_{gh},c_mb_n\mathrm{\Omega }_{gh}>=\theta (m)\delta _{m+n}`$ $`<\mathrm{\Omega }_{gh},b_mb_n\mathrm{\Omega }_{gh}>=0<\mathrm{\Omega }_{gh},c_mc_n\mathrm{\Omega }_{gh}>=0`$ (9.8) and we can compute the commutators from the statement using Wick theorem. $`\mathrm{}`$ Next we have ###### Corollary 9.2 Let us consider in the Hilbert spaces $`_{cov}^{[\mu ,\alpha ]}_1^{gh}`$ the following operators: $`L_m^{(\alpha )}`$ cf. (7.24) and $$_m^{(\alpha )}=L_m^{(\alpha )}I_2+I_1l_m^{(1)};$$ (9.9) then we have $`[_m^{(\alpha )},_n^{(\alpha )}]=(mn)_{m+n}^{(\alpha )}+m\left({\displaystyle \frac{D26}{12}}m^2+2a{\displaystyle \frac{D2}{12}}\right)\delta _{m+n}I`$ $`(_m^{(\alpha )})^{}=_m^{(\alpha )}.`$ (9.10) In this enlarged Hilbert space we have : ###### Proposition 9.3 The following operator $$QL_m^{(\alpha )}c_m\frac{1}{2}(mn):c_mc_nb_{m+n}:$$ (9.11) is well defined on the algebraic Hilbert space and is formally self-adjoint; it verifies $$Q^2=0$$ (9.12) iff $`D=26`$ and $`a=1`$. Proof: We separated the non-zero modes as before: $$Q=Q_0+_0^{(\alpha )}c_0+C_0^{(1)}b_0$$ (9.13) where $$Q_0\underset{m0}{}_m^{(\alpha )}c_m\frac{1}{2}\underset{m,n0}{}\underset{m+n0}{}(mn):c_mc_nb_{m+n}:$$ (9.14) and $$C_m^{(1)}\frac{1}{2}\underset{p+q=m}{}(pq):c_pc_q:=\frac{1}{2}\underset{p+q=m}{}(pq)c_pc_q$$ (9.15) The most convenient way to prove the theorem is the following. One proves by direct computation (using our preferred method based on Wick theorem) the following formulas: $`\{Q,b_m\}=_m^{(\alpha )}\{Q,c_m\}=C_m^{(1)}`$ $`[Q,_m^{(\alpha )}]=\rho _mc_m[Q,K_m]=m{\displaystyle K_{mn}c_n}`$ (9.16) where $$\rho _mm\left(\frac{D26}{12}m^2+2a\frac{D2}{12}\right).$$ (9.17) We now use the following observation. According to proposition 8.1 we can take in $``$ the following basis: $$\mathrm{\Psi }^{}=b_ic_jL_m^{(\alpha )}K_nf$$ (9.18) where $`f`$ are DDF states, the indices of type $`m,n`$ are strictly positive and the indices of the type $`i,j`$ are $`0`$. It is easy to substitute $`L_m^{(\alpha )}=_m^{(\alpha )}l_m^{(1)}`$ and consider the new basis $$\mathrm{\Psi }=b_ic_j_m^{(\alpha )}K_nf$$ (9.19) Because $`_m^{(\alpha )}f=0m0`$ we easily find out that $$Qf=0$$ (9.20) for any DDF state $`f`$. The operator $`Q`$ is perfectly well defined by (9.16) and (9.20); indeed we know how to act with $`Q`$ on states of the form (9.19): we commute $`Q`$ using (9.16) till it hits a DDF state and gives $`0`$ according to (9.20). Now it is easy to obtain from (9.16): $$\{Q^2,b_m\}=\rho _mb_m\{Q^2,c_m\}=0[Q^2,_m^{(\alpha )}]=\rho _mC_m^{(1)}[Q^2,K_m]=0.$$ (9.21) Because we obviously have $`Q^2f=0`$ it immediately follows that $$Q^2=0\rho _m=0D=26,a=1$$ (9.22) i.e. the statement of the theorem. $`\mathrm{}`$ ###### Remark 9.4 One can directly prove that $$Q^2=\frac{1}{2}m\left(\frac{D26}{12}m^2+2a\frac{D2}{12}\right):c_mc_m:$$ (9.23) which is another way to obtain the result. Let us also note that if the conditions $`D=26,a=1`$ are meet then we also have no anomalies in the Virasoro algebra: $$[_m^{(\alpha )},_n^{(\alpha )}]=(mn)_{m+n}^{(\alpha )}.$$ (9.24) To analyze the cohomology of the BRST operator $`Q`$ we need the following result: ###### Proposition 9.5 The operator $`\stackrel{~}{Q}`$ is well defined on the algebraic Hilbert space through the following formulas: $$\{\stackrel{~}{Q},b_m\}=0\{\stackrel{~}{Q},c_m\}=\delta _mI[\stackrel{~}{Q},_m^{(\alpha )}]=mb_m[\stackrel{~}{Q},K_m]=0$$ (9.25) and $$\stackrel{~}{Q}f=0$$ (9.26) for any DDF state $`f`$. We also have $$\stackrel{~}{Q}^{}=\stackrel{~}{Q}\stackrel{~}{Q}^2=0.$$ (9.27) Proof: Because the operators $`b_m,c_m,_m^{(\alpha )},K_m`$ are connected by various relations, we have to verify the Jacobi identities of the type: $$[[X,Y],\stackrel{~}{Q}]_{\mathrm{graded}}+\mathrm{cyclic}\mathrm{permutations}=0$$ (9.28) where $`X,Y`$ are operators from the set $`b_m,c_m,_m^{(\alpha )},K_m.`$ The non-trivial ones are corresponding to the pairs $`(_m^{(\alpha )},_n^{(\alpha )})`$ and $`(_m^{(\alpha )},c_n)`$ and are easily checked. $`\mathrm{}`$ The main result is the following ###### Theorem 9.6 If $`\mathrm{\Psi }`$ verifies $`Q\mathrm{\Psi }=0`$ then it is of the form $$\mathrm{\Psi }=Q\mathrm{\Phi }+f_1+b_0f_2+c_0f_3$$ (9.29) where $`f_j`$ are DDF states. Proof: A good strategy to determine the cohomology of the operator $`Q`$ is to mimic Hodge theorem i.e. to find a homotopy operator $`\stackrel{~}{Q}`$ such that the spectrum of the “Laplacian” $$\mathrm{\Delta }Q\stackrel{~}{Q}+\stackrel{~}{Q}Q$$ (9.30) can be easily determined. We take such an operator to be the one determined in the previous proposition. It is now elementary to see that the Laplace operator is alternatively given by: $`[\mathrm{\Delta },b_m]=mb_m[\mathrm{\Delta },c_m]=mc_m[\mathrm{\Delta },_m^{(\alpha )}]=m_m^{(\alpha )}[\mathrm{\Delta },K_m]=mK_m`$ $`\mathrm{\Delta }f=0.`$ (9.31) It follows that we have on states of the form (9.19) $$\mathrm{\Delta }\mathrm{\Psi }=(m+n+i+j)\mathrm{\Psi };$$ (9.32) we observe that the eigenvalues from (9.32) are $`0`$ the equality sign being true only for vectors of the form $`f_1+b_0f_2+c_0f_3`$ with $`f_j`$ DDF states. It means that every vector in $``$ is of the form $$\mathrm{\Psi }=\mathrm{\Psi }_0+f_1+b_0f_2+c_0f_3$$ (9.33) where $`\mathrm{\Psi }_0`$ belongs to the subspace $`^{}`$ of vectors with strictly positive eigenvalues of $`\mathrm{\Delta }.`$ Now suppose that the vector $`\mathrm{\Psi }`$ verifies the equation from the statement $`Q\mathrm{\Psi }=0.`$ Then it is easy to see that we also have $`Q\mathrm{\Psi }_0=0`$ so $`\mathrm{\Delta }\mathrm{\Psi }_0=Q\stackrel{~}{Q}\mathrm{\Psi }_0.`$ We can write $`\mathrm{\Psi }_0=\mathrm{\Psi }_\omega `$ where $`\mathrm{\Psi }_\omega `$ are linear independent vectors from $`^{}`$ corresponding to distinct eigenvalues $`\omega >0`$. Then the preceding relation is equivalent to $`\mathrm{\Psi }_\omega =\frac{1}{\omega }Q\stackrel{~}{Q}\mathrm{\Psi }_\omega ;`$ if we define $`\mathrm{\Phi }=\frac{1}{\omega }\stackrel{~}{Q}\mathrm{\Psi }_\omega `$ then it follows that $`\mathrm{\Psi }_0=Q\mathrm{\Phi }`$ and this finishes the proof. $`\mathrm{}`$ We conclude that the cohomology of the operator $`Q`$ is described by three copies of the physical space of DDF states. To eliminate this tripling we proceed as follows. We construct the Hilbert space $`_1^{gh}`$ such that it also verifies $$b_0\mathrm{\Omega }_{gh}=0;$$ (9.34) in this case $`_1^{gh}`$ is a Fock space. Then we construct $``$ as above and consider the subspace $$_0\{\mathrm{\Psi }|b_0\mathrm{\Psi }=0_0^{(\alpha )}\mathrm{\Psi }=0\}.$$ (9.35) This subspace is left invariant by the operator $`Q`$; if $`\mathrm{\Psi }_0`$ verifies $`Q\mathrm{\Psi }=0`$ then we have similarly with the preceding theorem $$\mathrm{\Psi }=Q\mathrm{\Phi }+f$$ (9.36) with $`f`$ some DDF state. ## 10 The Covariant Neveau-Schwarz Superstring We proceed as before from the Hilbert space $`_{\mathrm{cov}}^{(NS)}`$ generated by the operators $`\alpha _m^\mu ,m,\mu =0,\mathrm{},D`$ and $`b_r^\mu ,r\frac{1}{2}+,\mu =0,\mathrm{},D`$ such that: $`[\alpha _m^\mu ,\alpha _n^\nu ]=\eta _{\mu \nu }m\delta _{m+n}I,m,n`$ $`\alpha _m^\mu \mathrm{\Omega }=0,m>0`$ $`(\alpha _m^\mu )^+=\alpha _m^\mu m;`$ $`\{b_r^\mu ,b_s^\nu \}=\eta _{\mu \nu }m\delta _{r+s}I,r,s`$ $`b_r^\mu \mathrm{\Omega }=0,r>0`$ $`(b_r^\mu )^+=b_r^\mu r`$ (10.1) and define the covariant Virasoro operators $$L_m^{(NS)}\frac{1}{2}\eta _{\mu \nu }\underset{n}{}:\alpha _{mn}^\mu \alpha _n^\nu :\frac{1}{2}\eta _{\mu \nu }\underset{r1/2+}{}r:b_r^\mu b_{m+r}^\nu :a\delta _mI$$ (10.2) and the supersymmetric partners $$G_r\eta _{\mu \nu }\underset{n}{}\alpha _n^\mu b_{n+r}^\nu $$ (10.3) such that we have the following relations: $`[L_m^{(NS)},L_n^{(NS)}]=(mn)L_{m+n}^{(NS)}+\left[D{\displaystyle \frac{m(m^21)}{8}}+2ma\right]\delta _{m+n}I.`$ $`[L_m^{(NS)},G_r]=\left({\displaystyle \frac{m}{2}}r\right)G_{m+r}`$ $`\{G_r,G_s\}=2L_{r+s}^{(NS)}+\left[{\displaystyle \frac{D}{2}}\left(r^2{\displaystyle \frac{1}{4}}\right)+2a\right]\delta _{r+s}I`$ $`(L_m^{(NS)})^{}=L_m^{(NS)}G_r^{}=G_r.`$ (10.4) In this Hilbert space we have an action of the supersymmetric Poincaré algebra. Then we can obtain the Neveau-Schwarz case if we take $`D=10,a=1/2`$ and restrict the states by the conditions: $$L_m^{(NS)}\mathrm{\Psi }=0m0G_r\mathrm{\Psi }=0r>0.$$ (10.5) The DDF states can be constructed as before. First we have to construct operators $`A_m^j,B_r^j,j=1,\mathrm{},D1`$ such that they verify the same algebra as the operators $`\alpha _m^j,b_r^j`$ and they commute with $`G_r,r;`$ (this implies that they commute with $`L_m,m.)`$ The DDF states are generated by these operators from the vacuum. For the BRST description we need to enlarge the ghost space: we consider the Fock space $`_2^{gh}`$ generated by the operators $`\beta _r,\gamma _rr\frac{1}{2}+`$ from the vacuum $`\mathrm{\Omega }_{gh}_2^{gh}`$; we assume that $$\beta _r\mathrm{\Omega }_{gh}=0\gamma _r\mathrm{\Omega }_{gh}=0r>0.$$ (10.6) These operators are subject to the following anticommutation relations: $$[\beta _r,\beta _s]=0[\gamma _r,\gamma _s]=0[\gamma _r,b_s]=\delta _{r+s}I;$$ (10.7) we also suppose that there is a conjugation operation in $`_2^{gh}`$ such that $$\beta _r^{}=\beta _r\gamma _r^{}=\gamma _r.$$ (10.8) We can define as usual the algebraic Hilbert space (the subspace of vectors generated by a finite number of operators $`\beta _r,\gamma _rr0)`$ and normal ordering in $`_2^{gh}.`$ ###### Proposition 10.1 The following operators $$l_m^{(2)}=\underset{r1/2+}{}\left(\frac{m}{2}+r\right):\beta _{mr}\gamma _r:$$ (10.9) are well defined on the algebraic Hilbert space and are verifying: $`[l_m^{(2)},\beta _r]=\left({\displaystyle \frac{m}{2}}r\right)\beta _{m+r}[l_m^{(2)},\gamma _r]=\left({\displaystyle \frac{3m}{2}}+r\right)\gamma _{m+r}`$ $`[l_m^{(2)},l_n^{(2)}]=(mn)l_{m+n}^{(2)}+{\displaystyle \frac{1}{12}}m(11m^2+1)\delta _{m+n}I`$ $`(l_m^{(2)})^{}=l_m^{(2)}.`$ (10.10) Proof: The $`2`$-point functions are $`<\mathrm{\Omega }_{gh},\beta _r\gamma _s\mathrm{\Omega }_{gh}>=\theta (r)\delta _{r+s}<\mathrm{\Omega }_{gh},\gamma _r\beta _s\mathrm{\Omega }_{gh}>=\theta (r)\delta _{r+s}`$ $`<\mathrm{\Omega }_{gh},\beta _r\beta _s\mathrm{\Omega }_{gh}>=0<\mathrm{\Omega }_{gh},\gamma _r\gamma _s\mathrm{\Omega }_{gh}>=0`$ (10.11) and we can compute the commutators from the statement using Wick theorem. $`\mathrm{}`$ Next we have ###### Corollary 10.2 Let us consider in the Hilbert spaces $`_{NS}^{gh}_1^{gh}_2^{gh}`$ the following operators $$l_m^{(NS)}=l_m^{(1)}I_2+I_1l_m^{(2)};$$ (10.12) then we have $`[l_m^{(NS)},l_n^{(NS)}]=(mn)l_{m+n}^{(NS)}+{\displaystyle \frac{1}{4}}m(15m^2)\delta _{m+n}I`$ $`(l_m^{(NS)})^{}=l_m^{(NS)}.`$ (10.13) Next, we consider in $`_{NS}^{gh}`$ the following operators $$g_r2\underset{n}{}b_n\gamma _{n+r}+\underset{n}{}\left(\frac{n}{2}r\right)c_n\beta _{n+r}$$ (10.14) which are well-defined on the algebraic Fock space. We have ###### Proposition 10.3 The following relations are verified $`g_r^{}=g_r`$ $`[l_m^{(NS)},g_r]=\left({\displaystyle \frac{m}{2}}r\right)g_{m+r}`$ $`\{g_r,g_s\}=2l_{r+s}^{(NS)}+\left({\displaystyle \frac{1}{4}}5r^2\right)\delta _{r+s}I.`$ (10.15) The proofs of the first two relations are elementary. For the last one we use Wick theorem. Next we have ###### Corollary 10.4 Let us consider in the Hilbert spaces $`_{cov}^{(b)}^s_{NS}^{gh}`$ where the skew tensor product $`^s`$ is such that we have normal (anti)commutation relations i.e. the Fermionic operators $`b_r^\mu `$ are anticommuting with $`b_m,c_m.`$ We then define the operators $`_m^{(NS)}=L_m^{(NS)}I_2+I_1l_m^{(NS)}`$ $`𝒢_r=G_rI_2+I_1g_r`$ (10.16) and we have $`[_m^{(NS)},_n^{(NS)}]=(mn)_{m+n}^{(NS)}+m\left({\displaystyle \frac{D10}{8}}m^2+2a{\displaystyle \frac{D2}{8}}\right)\delta _{m+n}I`$ $`\{𝒢_r,𝒢_s\}=2_{r+s}^{(NS)}+\left({\displaystyle \frac{D10}{8}}r^2+2a{\displaystyle \frac{D2}{8}}\right)\delta _{r+s}I`$ $`(_m^{(NS)})^{}=_m^{(NS)}𝒢_r^{}=𝒢_r.`$ (10.17) The anomalies cancel iff $`D=10,a=1/2`$. In this enlarged Hilbert space we have: ###### Proposition 10.5 The following operator $`Q=Q^{(NS)}{\displaystyle L_m^{(NS)}c_m}{\displaystyle \frac{1}{2}}{\displaystyle (mn)}:c_mc_nb_{m+n}:`$ $`+{\displaystyle G_r\gamma _r}+{\displaystyle c_ml_m^{(2)}}+{\displaystyle b_mC_m^{(2)}}`$ (10.18) where $$C_m^{(2)}\underset{r+s=m}{}:\gamma _r\gamma _s:=\underset{r+s=m}{}\gamma _r\gamma _s$$ (10.19) is well defined on the algebraic Hilbert space and it is formally self-adjoint; it verifies $$Q^2=0$$ (10.20) iff $`D=10`$ and $`a=1/2`$. Proof: It is convenient to denote by $`Q_j,j=1,\mathrm{},5`$ the five terms in the expression of the BRST charge and write $$Q=Q^{}+\underset{j=3}{\overset{5}{}}Q_j$$ (10.21) where the sum of the first two terms $`Q^{}`$ can be obtained from $`Q`$ of the preceding Section with the substitution $`L_m^{(\alpha )}L_m^{(NS)}`$ so we can use some of the computations performed there. We introduce the notation $$H_rk_\mu b_r^\mu .$$ (10.22) and we have as before: $`\{Q,b_m\}=_m^{(NS)}\{Q,c_m\}=C_m^{(NS)}C_m^{(1)}+C_m^{(2)}`$ $`[Q,\beta _r]=𝒢_r[Q,\gamma _r\}={\displaystyle \left(\frac{3m}{2}+r\right)c_m\gamma _{m+r}}`$ $`[Q,_m^{(NS)}]=\rho _mc_m\{Q,𝒢_r\}=\lambda _r\gamma _r`$ $`[Q,K_m]=m{\displaystyle K_{mn}c_n}+{\displaystyle H_{mr}\gamma _r}`$ $`\{Q,H_r\}={\displaystyle \left(\frac{m}{2}r\right)H_{rm}c_m}+{\displaystyle K_{rs}\gamma _s}.`$ (10.23) where $$\rho _mm\left(\frac{D10}{8}m^2+2a\frac{D2}{8}\right)\lambda _r\frac{D10}{2}r^2+2a\frac{D2}{8}.$$ (10.24) The only really complicated computation is for the anticommutator $`\{Q,𝒢_r\}.`$ One can prove that we can take in $``$ the following basis: $$\mathrm{\Psi }=b_ic_j\beta _r\gamma _s_m^{(NS)}𝒢_tK_nH_uf$$ (10.25) where $`f`$ are DDF states, the indices of the type $`m,n`$,$`r,s,t,u1/2+`$ are taking positive values and the indices of the type $`i,j`$ are $`0`$. Because $`L_m^{(NS)}f=0m0G_rf=0r>0`$ we easily find out that $$Qf=0$$ (10.26) for any DDF state $`f`$. We argue as before that the operator $`Q`$ is well defined by (10.23) and (10.26). Now it is easy to obtain from (10.23): $`[Q^2,b_m]=\rho _mc_m[Q^2,c_m]=0[Q^2,\beta _r]=\lambda _r\gamma _r[Q^2,\gamma _r\}=0`$ $`[Q^2,_m^{(NS)}]=\rho _mC_m^{(NS)}[Q^2,𝒢_r]=\lambda _rC_r^{(3)}[Q^2,K_m]=0[Q^2,H_r]=0.`$ (10.27) Because we obviously have $`Q^2f=0`$ it immediately follows that $$Q^2=0\rho _m=0\lambda _r=0D=10,a=1/2$$ (10.28) i.e. the statement of the theorem. $`\mathrm{}`$ To analyze the cohomology of the BRST operator $`Q`$ we construct as before, its homotopy: ###### Proposition 10.6 The operator $`\stackrel{~}{Q}`$ is well defined on the algebraic Hilbert space through the following formulas: $`\{\stackrel{~}{Q},b_m\}=0\{\stackrel{~}{Q},c_m\}=\delta _mI\{\stackrel{~}{Q},\beta _r\}=0\{\stackrel{~}{Q},\gamma _r\}=0`$ $`[\stackrel{~}{Q},_m^{(NS)}]=mb_m\{\stackrel{~}{Q},𝒢_r\}=r\beta _r[\stackrel{~}{Q},K_m]=0[\stackrel{~}{Q},H_r]=0`$ (10.29) and $$\stackrel{~}{Q}f=0$$ (10.30) for any DDF state $`f`$. We also have $$\stackrel{~}{Q}^{}=\stackrel{~}{Q}\stackrel{~}{Q}^2=0.$$ (10.31) Proof: We have to verify the Jacobi identities of the type: $$[[X,Y],\stackrel{~}{Q}]_{\mathrm{graded}}+\mathrm{cyclic}\mathrm{permutations}=0$$ (10.32) where $`X,Y`$ are operators from the set $`b_m,c_m,\beta _r,\gamma _r,_m^{(NS)},𝒢_r,K_m,H_r.`$ We have some non-trivial ones corresponding to pairs $`(_m^{(NS)},_n^{(NS)}),(_m^{(NS)},c_n),(_m^{(NS)},𝒢_r),(𝒢_r,𝒢_s)`$ and $`(𝒢_r,\gamma _s).`$ $`\mathrm{}`$ The main result is similar to the one in the previous Section: ###### Theorem 10.7 If $`\mathrm{\Psi }`$ verifies $`Q\mathrm{\Psi }=0`$ then it is of the form $$\mathrm{\Psi }=Q\mathrm{\Phi }+f_1+b_0f_2+c_0f_3$$ (10.33) where $`f_j`$ are DDF states. Proof: The “Laplacian” is as before $`\mathrm{\Delta }Q\stackrel{~}{Q}+\stackrel{~}{Q}Q`$ It is now elementary to determine the alternative expression: $`[\mathrm{\Delta },b_m]=mb_m[\mathrm{\Delta }^{(NS)},c_m]=mc_m[\mathrm{\Delta },\beta _r]=r\beta _r[\mathrm{\Delta },\gamma _r]=r\gamma _r`$ $`[\mathrm{\Delta },_m^{(NS)}]=m_m^{(NS)}[\mathrm{\Delta },𝒢_r]=r𝒢_r[\mathrm{\Delta },K_m]=mK_m[\mathrm{\Delta },H_r]=rH_r`$ $`\mathrm{\Delta }f=0.`$ (10.34) It follows that we have $$\mathrm{\Delta }\mathrm{\Psi }=(m+n+i+j+r+s+t+u)\mathrm{\Psi };$$ (10.35) we observe that the eigenvalues from (9.32) are $`0`$ the equality sign being true only for vectors of the form $`f_1+b_0f_2+c_0f_3`$ with $`f_j`$ DDF states, as in the previous Section. From now on the argument is the same as there. $`\mathrm{}`$ To eliminate this tripling we proceed, again, as in the previous Section: we construct the Fock space $`_{NS}^{gh}`$ such that it also verifies $$b_0\mathrm{\Omega }_{gh}=0.$$ (10.36) Then we construct $``$ as above and consider the subspace $$_0\{\mathrm{\Psi }|b_0\mathrm{\Psi }=0_0^{(NS)}\mathrm{\Psi }=0\}.$$ (10.37) This subspace is left invariant by the operator $`Q`$ and if $`Q^{(NS)}\mathrm{\Psi }=0`$ then we have similarly with the preceding theorem $$\mathrm{\Psi }=Q^{(NS)}\mathrm{\Phi }+f$$ (10.38) with $`f`$ some DDF state. ## 11 The Quantum Ramond Superstring The modification with respect to the preceding Section are minimal. We start from the Hilbert space $`_{\mathrm{cov}}^{(R)}`$ generated by the operators $`\alpha _m^\mu ,m,\mu =0,\mathrm{},D`$ and $`d_m^\mu ,m,\mu =0,\mathrm{},D`$ such that: $`[\alpha _m^\mu ,\alpha _n^\nu ]=\eta _{\mu \nu }m\delta _{m+n}I,m,n`$ $`\alpha _m^\mu \mathrm{\Omega }=0,m>0`$ $`(\alpha _m^\mu )^+=\alpha _m^\mu m;`$ $`\{d_m^\mu ,d_n^\nu \}=\eta _{\mu \nu }m\delta _{m+n}I,m,n`$ $`d_m^\mu \mathrm{\Omega }=0,m>0`$ $`(d_m^\mu )^+=d_m^\mu m`$ (11.1) and define the covariant Virasoro operators $$L_m^{(R)}\frac{1}{2}\eta _{\mu \nu }\underset{n}{}:\alpha _{mn}^\mu \alpha _n^\nu :\frac{1}{2}\eta _{\mu \nu }\underset{n}{}n:d_n^\mu d_{m+n}^\nu :a\delta _mI$$ (11.2) and the supersymmetric partners $$F_m\eta _{\mu \nu }\underset{n}{}\alpha _n^\mu d_{n+m}^\nu $$ (11.3) such that we have the following relations: $`[L_m^{(R)},L_n^{(R)}]=(mn)L_{m+n}^{(R)}+m\left({\displaystyle \frac{D}{8}}m^2+2ma\right)\delta _{m+n}I.`$ $`[L_m^{(R)},F_n]=\left({\displaystyle \frac{m}{2}}n\right)F_{m+n}`$ $`\{F_m,F_n\}=2L_{m+n}^{(R)}+\left({\displaystyle \frac{D}{2}}m^2+2a\right)\delta _{m+n}I`$ $`(L_m^{(R)})^{}=L_m^{(R)}F_m^{}=F_m.`$ (11.4) In this Hilbert space we have an action of the supersymmetric Poincaré algebra. Then we can obtain the Ramond case if we take $`D=10,a=0`$ and restrict the states by the conditions: $$L_m^{(R)}\mathrm{\Psi }=0m0F_m\mathrm{\Psi }=0m0.$$ (11.5) The DDF states can be constructed as before. First we have to construct operators $`A_m^j,D_m^j,j=1,\mathrm{},D1`$ such that they verify the same algebra as the operators $`\alpha _m^j,d_m^j`$ and commute with $`F_m,m`$ (this implies that they commute with $`L_m,m.)`$ The DDF states are generated by these operators from the vacuum. For the BRST description we need to enlarge the ghost space: we consider the Fock space $`_3^{gh}`$ generated by the operators $`\beta _m,\gamma _mm`$ from the vacuum $`\mathrm{\Omega }_{gh}_3^{gh}`$; we assume that $$\beta _m\mathrm{\Omega }_{gh}=0\gamma _m\mathrm{\Omega }_{gh}=0m>0.$$ (11.6) These operators are subject to the following anticommutation relations: $$[\beta _m,\beta _n]=0[\gamma _m,\gamma _n]=0[\gamma _m,b_n]=\delta _{m+n}I;$$ (11.7) we also suppose that there is a conjugation operation in $`_3^{gh}`$ such that $$\beta _m^{}=\beta _m\gamma _m^{}=\gamma _m.$$ (11.8) We can define as usual the algebraic Hilbert space (the subspace of vectors generated by a finite number of operators $`\beta _m,\gamma _mm0)`$ and normal ordering in $`_m^{gh}.`$ ###### Proposition 11.1 The following operators $$l_m^{(3)}=\underset{n}{}\left(\frac{m}{2}+n\right):\beta _{mn}\gamma _n:$$ (11.9) are well defined on the algebraic Hilbert space and are verifying: $`[l_m^{(3)},\beta _n]=\left({\displaystyle \frac{m}{2}}n\right)\beta _{m+n}[l_m^{(3)},\gamma _n]=\left({\displaystyle \frac{3m}{2}}+n\right)\gamma _{m+n}`$ $`[l_m^{(3)},l_n^{(3)}]=(mn)l_{m+n}^{(3)}+{\displaystyle \frac{1}{12}}m(11m^22)\delta _{m+n}I`$ $`(l_m^{(3)})^{}=l_m^{(3)}.`$ (11.10) Proof: We proceed as in Section 9: first we split $$l_m^{(3)}=\stackrel{~}{l}_m^{}+\frac{m}{2}\beta _m\gamma _0+\frac{3m}{2}\beta _0\gamma _m$$ (11.11) where the first term includes only the non-zero modes. For these modes the $`2`$-point functions are $`<\mathrm{\Omega }_{gh},\beta _m\gamma _n\mathrm{\Omega }_{gh}>=\theta (m)\delta _{m+n}<\mathrm{\Omega }_{gh},\gamma _m\beta _n\mathrm{\Omega }_{gh}>=\theta (m)\delta _{m+n}`$ $`<\mathrm{\Omega }_{gh},\beta _m\beta _n\mathrm{\Omega }_{gh}>=0<\mathrm{\Omega }_{gh},\gamma _m\gamma _n\mathrm{\Omega }_{gh}>=0`$ (11.12) and we can compute the commutators from the statement using Wick theorem. $`\mathrm{}`$ Next we have ###### Corollary 11.2 Let us consider in the Hilbert spaces $`_R^{gh}_1^{gh}_3^{gh}`$ the following operators $$l_m^{(R)}=l_m^{(1)}I_2+I_1l_m^{(3)};$$ (11.13) then we have $`[l_m^{(R)},l_n^{(R)}]=(mn)l_{m+n}^{(NS)}{\displaystyle \frac{5}{4}}m^3\delta _{m+n}I`$ $`(l_m^{(R)})^{}=l_m^{(R)}.`$ (11.14) Next, we define in $`_R^{gh}`$ the following operators $$f_m2\underset{n}{}b_n\gamma _{n+m}+\underset{n}{}\left(\frac{n}{2}m\right)c_n\beta _{n+m}$$ (11.15) which are well-defined on the algebraic Fock space. We have ###### Proposition 11.3 The following relations are verified $`f_m^{}=f_m`$ $`[l_m^{(R)},f_n]=\left({\displaystyle \frac{m}{2}}n\right)f_{m+n}`$ $`\{f_m,f_n\}=2l_{m+n}^{(R)}5m^2\delta _{m+n}I.`$ (11.16) The proofs of the first two relations are elementary. For the last one we use Wick theorem. Next we have ###### Corollary 11.4 Let us consider in the Hilbert spaces $`_{cov}^{(b)}^s_R^{gh}`$ where the skew tensor product $`^s`$ is such that we have normal (anti)commutation relations i.e. the Fermionic operators $`d_m^\mu `$ are anticommuting with $`b_m,c_m.`$ We then define the following operators $`_m^{(R)}=L_m^{(R)}I_2+I_1l_m^{(R)}`$ $`_m=F_mI_2+I_1f_m`$ (11.17) then we have $`[_m^{(R)},_n^{(R)}]=(mn)_{m+n}^{(R)}+m\left({\displaystyle \frac{D10}{8}}m^2+2a\right)\delta _{m+n}I`$ $`[_m^{(R)},_n]=\left({\displaystyle \frac{m}{2}}n\right)_{m+n}`$ $`\{_m,_n\}=2_{m+n}^{(R)}+\left({\displaystyle \frac{D10}{8}}m^2+2a\right)\delta _{r+s}I`$ $`(_m^{(R)})^{}=_m^{(R)}_m^{}=_m.`$ (11.18) The anomalies cancel iff $`D=10,a=0`$. In this enlarged Hilbert space we have: ###### Proposition 11.5 The following operator $`Q=Q^{(R)}{\displaystyle L_m^{(R)}c_m}{\displaystyle \frac{1}{2}}{\displaystyle (mn)}:c_mc_nb_{m+n}:`$ $`+{\displaystyle F_m\gamma _m}+{\displaystyle c_ml_m^{(3)}}+{\displaystyle b_mC_m^{(4)}}`$ (11.19) where $$C_m^{(4)}\underset{p+q=m}{}:\gamma _p\gamma _q:=\underset{p+q=m}{}\gamma _p\gamma _q$$ (11.20) is well defined on the algebraic Hilbert space and is formally self-adjoint; it verifies $$Q^2=0$$ (11.21) iff $`D=10`$ and $`a=0`$. Proof: It is convenient to denote by $`Q_j,j=1,\mathrm{},5`$ the five terms in the expression of the BRST charge and write $$Q=Q^{}+\underset{j=3}{\overset{5}{}}Q_j$$ (11.22) where the sum of the first two terms $`Q^{}`$ can be obtained from $`Q`$ of the preceding Section with the substitution $`L_m^{(\alpha )}L_m^{(R)}`$ so we can use some of the computations performed there. We introduce the notation $$H_mk_\mu d_m^\mu .$$ (11.23) and we have as before: $`\{Q,b_m\}=_m^{(R)}\{Q,c_m\}=C_m^{(R)}C_m^{(1)}+C_m^{(4)}`$ $`[Q,\beta _m]=_m[Q,\gamma _m\}={\displaystyle \left(\frac{3n}{2}+m\right)c_n\gamma _{m+n}}`$ $`[Q,_m^{(R)}]=\rho _mc_m\{Q^{(R)},_m\}=\lambda _m\gamma _m`$ $`[Q,K_m]=m{\displaystyle K_{mn}c_n}+{\displaystyle H_{mn}\gamma _n}`$ $`\{Q,H_m\}={\displaystyle \left(\frac{n}{2}m\right)H_{mn}c_m}+{\displaystyle K_{mn}\gamma _n}.`$ (11.24) where $$\rho _mm\left(\frac{D10}{8}m^2+2a\right)\lambda _m\frac{D10}{2}m^2+2a.$$ (11.25) One can prove that we can take in $``$ the following basis: $$\mathrm{\Psi }=b_ic_j\beta _p\gamma _q_m^{(NS)}_lK_nH_kf$$ (11.26) where $`f`$ are DDF states, the indices of the type $`m,n`$ are taking positive values and the indices of the type $`i,j,p,q,k,l`$ are $`0`$. Because $`L_m^{(R)}f=0_mf=0m0`$ we easily find out that $$Qf=0$$ (11.27) for any DDF state $`f`$. We argue as before that the operator $`Q`$ is well defined by (11.24) and (11.27). Now it is easy to obtain from (11.24) that $`Q^2`$ commutes with all the operators from (11.26). Because we have $`Q^2f=0`$ it immediately follows that $$Q^2=0\rho _m^{(R)}=0\lambda _m^{(R)}=0D=10,a=0$$ (11.28) i.e. the statement of the theorem. $`\mathrm{}`$ To analyze the cohomology of the BRST operator $`Q`$ we construct its homotopy: ###### Proposition 11.6 The operator $`\stackrel{~}{Q}`$ is well defined on the algebraic Hilbert space through the following formulas: $`\{\stackrel{~}{Q},b_m\}=0\{\stackrel{~}{Q},c_m\}=\delta _mI\{\stackrel{~}{Q},\beta _m\}=0`$ $`\{\stackrel{~}{Q},\gamma _m\}=0[\stackrel{~}{Q},_m^{(R)}]=mb_m\{\stackrel{~}{Q},_m\}=m\beta _m`$ $`[\stackrel{~}{Q},K_m]=0[\stackrel{~}{Q},H_m]=0`$ (11.29) and $$\stackrel{~}{Q}f=0$$ (11.30) for any DDF state $`f`$. We also have $$\stackrel{~}{Q}^{}=\stackrel{~}{Q}\stackrel{~}{Q}^2=0.$$ (11.31) Proof: We have to verify the Jacobi identities of the type: $$[[X,Y],\stackrel{~}{Q}]_{\mathrm{graded}}+\mathrm{cyclic}\mathrm{permutations}=0$$ (11.32) where $`X,Y`$ are operators from the set $`b_m,c_m,\beta _m,\gamma _m,_m^{(R)},_m,K_m,H_m.`$ We have some non-trivial ones corresponding to pairs $`(_m^{(R)},_n^{(R)}),(_m^{(R)},c_n),(_m^{(R)},_r),(_m,_n)`$ and $`(_m,\gamma _m).`$ $`\mathrm{}`$ The main result is similar to the one in the previous Section. However, the degeneracy is infinite in this case, so to avoid this problem we work directly in a smaller Hibert space: we construct the Fock space $`_R^{gh}`$ such that it also verifies $$b_0\mathrm{\Omega }_{gh}=0\beta _0\mathrm{\Omega }_{gh}=0.$$ (11.33) Then we construct $``$ as above and consider the subspace $$_0\{\mathrm{\Psi }|b_0\mathrm{\Psi }=0\beta _0\mathrm{\Psi }=0_0^{(R)}\mathrm{\Psi }=0_0\mathrm{\Psi }=0\}.$$ (11.34) This subspace is left invariant by the operator $`Q^{(R)}`$ and we have the following result: ###### Theorem 11.7 If $`\mathrm{\Psi }_0`$ verifies $`Q\mathrm{\Psi }=0`$ then it is of the form $$\mathrm{\Psi }=Q\mathrm{\Phi }+f$$ (11.35) where $`f`$ is a DDF state and $`\mathrm{\Phi }_0.`$ Proof: First we have to prove that the states from $`_0`$ are obtained applying the operators $`b_m,c_m,\beta _m,\gamma _m,_m^{(R)},_m,K_m,H_m`$ with $`m>0`$. The “Laplacian” is as before $`\mathrm{\Delta }Q\stackrel{~}{Q}+\stackrel{~}{Q}Q`$ It is now elementary to determine that the Laplace operator is alternatively given by: $`[\mathrm{\Delta },b_m]=mb_m[\mathrm{\Delta },c_m]=mc_m[\mathrm{\Delta },\beta _m]=m\beta _m[\mathrm{\Delta },\gamma _m]=m\gamma _m`$ $`[\mathrm{\Delta },_m^{(R)}]=m_m^{(R)}[\mathrm{\Delta },_m]=m_m[\mathrm{\Delta },K_m]=mK_m[\mathrm{\Delta },H_m]=mH_m`$ $`\mathrm{\Delta }f=0.`$ (11.36) It follows that we have $$\mathrm{\Delta }\mathrm{\Psi }=(m+n+i+j+p+q+k+l)\mathrm{\Psi };$$ (11.37) we observe that the eigenvalues from (9.32) are $`0`$ the equality sign being true only for DDF states. From now on the argument is the same as before. $`\mathrm{}`$ ## 12 Conclusions The main results of this paper are: a) An elementary treatment of the quantum string models relying only on Wick theorem and paying attention to the domain problems. b) A derivation of the DDF operators without using vertex algebras. c) The clarification of the equivalence between the light-cone and covariant formalism using standard results in induced representation theory; this point seems to be missing from the literature. d) An elementary derivation of the BRST cohomology. A comparison with the standard literature is useful on this point: In one uses a basis of the type (9.19): $$\mathrm{\Psi }_{I,J,M,N}=b_{i_1}\mathrm{}b_{i_\beta }c_{j_1}\mathrm{}c_{j_\gamma }_{m_1}\mathrm{}_{m_\lambda }K_{n_1}\mathrm{}K_{n_\kappa }f_{I,J,M,N}$$ (12.1) where it can be arranged such that the DDF states $`f_{I,J,M,N}`$ are completely symmetric in the indices $`M=\{m_1,\mathrm{},m_\lambda \}`$ and in the indices $`N=\{n_1,\mathrm{},n_\kappa \};`$ of course we have complete antisymmetry in the indices $`I=\{i_1,\mathrm{},i_\beta \}`$ and in the indices $`J=\{j_1,\mathrm{},j_\gamma \}.`$ Then one decomposes the f’s according to Young diagrams (separately for $`IM`$ and $`JN).`$ We have in both cases only two projectors: one piece is eliminated by the condition $`Q\mathrm{\Psi }=0`$ and the other one can be proved to by a coboundary up to states of the form $`f_1+b_0f_2.`$ In there are two proofs, one based on a similar idea of Hodge theory (however the expression of the Laplacian seems to be different and the spectral analysis is not provided) and the other proof relies on the use of spectral sequences. The proof from , makes a convenient rescaling by a parameter $`\beta `$ and assumes that the states $`\mathrm{\Psi }(\beta )`$ are polynomials in this parameter which is an unjustified restriction. The proof from is closely related and assumes that a certain infinite series is convergent. The proof from relies on the existence of the operators $`D_n`$ formally given by: $`:\left(1K_mz^m\right)^1:=z^nD_n;`$ such operators are also used in the construction of the DDF states for the superstring models.
warning/0506/quant-ph0506116.html
ar5iv
text
# Efficient optical quantum information processing ## 1 Introduction It has been known for a number of years that processing information quantum mechanically enables certain communication and computation tasks that cannot be performed with conventional Information Technology (IT). The list of applications continues to expand, and there are extensive experimental efforts in many fields to realise the necessary building blocks for Quantum Information Processing (QIP) devices. One very appealing route (certainly in the short term when a pragmatic focus is on few-qubit applications) is that of optical QIP. In particular for quantum communication the qubits of choice are optical systems, since they can span long distances with minimal decoherence. In order to circumvent interconversion of the qubit species we need to process the optical qubits using optical circuits. Optical QIP is currently a very active research area, both theoretically and experimentally. The work of Knill, Laflamme and Milburn (KLM) has shown that in principle universal quantum computation is possible with linear optics , and there have been a number of recent experimental demonstrations of its gate components . However, due to the probabilistic nature of gates in linear optical QIP, it is practically rather inefficient (in terms of photon resources) to implement . Strong Kerr non-linearities are able to effectively mediate an interaction directly between photonic qubits. This would realise deterministic quantum gates and thus efficient optical QIP. In practice, however, such non-linearities are not available. On the other hand, much smaller non-linearities can be generated, for example, with electromagnetically induced transparencies (EIT). In this paper, we show that with modest additional optical resources these small non-linearities provide the building blocks for efficient optical QIP . We present a highly efficient non-absorbing single photon number resolving detector, a nondestructive two qubit parity detector, a nondestructive Bell state measurement, and a near deterministic controlled-not (CNOT) gate. ## 2 Quantum non-demolition detectors Before we discuss the construction of efficient quantum gates using weak non-linearities, let us first review the construction of a photon number quantum non-demolition (QND) measurement using a cross-Kerr non-linearity . The cross-Kerr non-linearity has a Hamiltonian of the form: $`H_{QND}=\mathrm{}\chi a^{}ac^{}c`$ (1) where the signal (probe) mode has the creation and annihilation operators given by $`a^{},a`$ ($`c^{},c`$) respectively and $`\chi `$ is the strength of the non-linearity. If the signal field contains $`n_a`$ photons and the probe field is in an initial coherent state with amplitude $`\alpha _c`$, the cross-Kerr non-linearity causes the combined system to evolve as $`|\mathrm{\Psi }(t)_{out}`$ $`=`$ $`e^{i\chi ta^{}ac^{}c}|n_a|\alpha _c=|n_a|\alpha _ce^{in_a\theta }.`$ (2) where $`\theta =\chi t`$ with $`t`$ being the interaction time for the signal and probe modes with the non-linear material. The Fock state $`|n_a`$ is unaffected by the interaction with the cross-Kerr non-linearity but the coherent state $`|\alpha _c`$ picks up a phase shift directly proportional to the number of photons $`n_a`$ in the signal $`|n_a`$ state. If we could measure this phase shift we could then infer the number of photons in the signal mode $`a`$. This can be achieved simply with a homodyne measurement (depicted schematically in figure (1)). The homodyne apparatus allows measurement of the quadrature operator $`x(\varphi )=ce^{i\varphi }+c^{}e^{i\varphi }`$ with expectation value $`x(\varphi )`$ $`=`$ $`2\mathrm{R}\mathrm{e}\left[\alpha _c\right]\mathrm{cos}\delta +i2\mathrm{I}\mathrm{m}\left[\alpha _c\right]\mathrm{sin}\delta `$ (3) where $`\delta =\varphi +n_a\theta `$ and $`\varphi `$ the phase of the local oscillator. For a real initial $`\alpha _c`$, a highly efficient homodyne measurement of the position $`X=a+a^{}`$ or momentum $`iY=aa^{}`$ quadratures yield the expectation values $`X`$ $`=`$ $`2\alpha _c\mathrm{cos}\left(n_a\theta \right)Y=2\alpha _c\mathrm{sin}\left(n_a\theta \right)`$ (4) with a unit variance. For the momentum quadrature this gives a signal-to-noise ratio $`\mathrm{SNR}_Y=2\alpha _c\mathrm{sin}\left(n_a\theta \right)`$ which should be much greater than unity for the different $`n_a`$ inputs to be distinguished. In more detail, if the inputs in mode $`a`$ are the Fock state $`|0`$ or $`|1`$, the respective outputs of the probe mode $`c`$ are the coherent states $`|\alpha _c`$ or $`|\alpha _ce^{i\theta }`$. The probability of misidentifying these states is given by $`P_{\mathrm{error}}={\displaystyle \frac{1}{2}}\mathrm{Erfc}\left[\alpha _c\mathrm{sin}\theta /\sqrt{2}\right]={\displaystyle \frac{1}{2}}\mathrm{Erfc}\left[\mathrm{SNR}_Y/2\sqrt{2}\right].`$ (5) A signal to noise ratio of $`\mathrm{SNR}_Y=6`$ would thus give $`P_{\mathrm{error}}10^3`$. To achieve the necessary phase shift we require $`\alpha _c\mathrm{sin}\theta 3`$ which can be achieved with a small non-linearity $`\theta `$ as long as the probe beam is intense enough. These results so far indicate that with a weak cross-Kerr non-linearity it is possible to build a high efficiency photon number resolving detector that does not absorb the photon from the signal mode. In reference we discuss an example of how this level of non-linearity could be achieved through electromagnetically induced transparency (EIT), and give some details for this approach with NV-diamond systems. In many optical quantum computation tasks our information is not encoded in photon number but polarization instead. When our information encoding is polarization based there are two separate detection tasks that we need to be able to perform. The first and simplest is just to determine for instance whether the polarization is in one of the basis states $`|H`$ or $`|V`$. This can be achieved by converting the polarization information to “which path” information on a polarizing beam-splitter. The “which path” information is photon number encoded in each path and hence a QND photon number measurement of each path will determine which polarization basis state the photon was originally in. The second task (one that is critically important for error correction codes in optics) is to determine whether our single-photon polarization-encoded qubit is present or not. That is, for the optical field under consideration we want to determine whether it contains a photon or not. If it does contain a photon, we do not want to destroy the information in its polarization state. This can be achieved by first converting the polarization qubit to a “which path” qubit. Each path then interacts with a weak cross-Kerr non-linearity $`\theta `$ with the same shared probe beam (Figure 2). If the photon is present in either path of this signal beam it induces a phase shift $`\theta `$ on the probe beam; however, with this configuration it is not possible to determine which path induced the phase shift. This allows the preservation of the “which path” and hence polarization information, ## 3 A Two Qubit Parity Gate Now that we have discussed the basic operation of a single photon quantum non-demolition detector, it is worthwhile asking whether this detection concept can be applied to several qubits. Basically if we want to perform a more “generalized” type of measurement between different photonic qubits, we could delay the homodyne measurement, instead having the probe beam interact with several cross-Kerr non-linearities where the signal mode is different in each case. The different signal modes could be from separate photonic qubits. The probe beam measurement then occurs after all these interactions in a collective way which could for instance allow a nondestructive detection that distinguishes superpositions and mixtures of the states $`|HH`$ and $`|VV`$ from $`|HV`$ and $`|VH`$. The key here is that we could have no net phase shift on the $`|HH`$ and $`|VV`$ terms while having a phase shift on the $`|HV`$ and $`|VH`$ terms. We will call this generalization a two qubit polarization parity QND gate. Let us now discuss the operation of this parity QND gate. Consider a general two qubit state which can be written as $`|\mathrm{\Psi }_{12}=\beta _0|HH+\beta _1|HV+\beta _2|VH+\beta _3|VV`$. This may be separable or entangled depending on the choices of $`\beta _i`$. As shown in Figure (3), these qubits are individually split on polarizing beam-splitters (PBS) into spatial encoded qubits which then interact with separate weak cross-Kerr non-linearities. The action of the PBS’s and cross-Kerr non-linearities evolves the combined system of photonic qubits and probe beam to $`|\psi _T`$ $`=`$ $`\left[\beta _0|HH+\beta _3|VV\right]|\alpha _c_p+\beta _1|HV|\alpha _ce^{i\theta }_p+\beta _2|VH|\alpha _ce^{i\theta }_p`$ It is now obvious that the $`|HH`$ and $`|VV`$ terms pick up no phase shift and remain coherent with respect to each other while the $`|HV`$ and $`|VH`$ pick up opposite sign phase shift $`\theta `$ which could allow them to be distinguished by a general homodyne/heterodyne measurement. We thus need to perform a measurement that does not allow the sign of the phase shift to be determined. With $`\alpha _c`$ real an $`X`$ homodyne measurement achieves this by projecting the probe beam to the position quadrature eigenstate $`|XX|`$. The resulting two photonic qubit state is then $`|\psi _X_T=f(X,\alpha _c)\left[\beta _0|HH+\beta _3|VV\right]`$ (6) $`+f(X,\alpha _c\mathrm{cos}\theta )\left[\beta _1e^{i\varphi (X)}|HV+\beta _2e^{i\varphi (X)}|VH\right]`$ where $`f(x,\beta )`$ $`=`$ $`\mathrm{exp}\left[{\displaystyle \frac{1}{4}}\left(x2\beta \right)^2\right]/(2\pi )^{1/4}`$ (7) $`\varphi (X)`$ $`=`$ $`\alpha _c\mathrm{sin}\theta (x2\alpha _c\mathrm{cos}\theta )\mathrm{mod2}\pi .`$ (8) We observe that $`f(X,\alpha )`$ and $`f(X,\alpha \mathrm{cos}\theta )`$ are two Gaussian curves with the mid point between the peaks located at $`X_0=\alpha _c\left[1+\mathrm{cos}\theta \right]`$ and the peaks separated by a distance $`X_d=2\alpha _c\left[1\mathrm{cos}\theta \right]`$. As long as this difference is large $`X_d\alpha _c\theta ^21`$, then there is little overlap between these curves. For an $`X`$ homodyne $`X>X_0`$ our solution (6) collapses to $`|\psi _{X>X_0}_T\beta _0|HH+\beta _3|VV`$ (9) while for $`X<X_0`$ we have $`|\psi _{X<X_0}_T\beta _1e^{i\varphi (X)}|HV+\beta _2e^{i\varphi (X)}|VH`$ (10) The action of this two mode polarization non-demolition parity gate is clear: It splits the even parity terms (9) nearly deterministically from the odd parity cases (10). Above we have chosen to call the even parity state {$`|HH,|VV`$} and the odd parity states {$`|HV,|VH`$}, but this is an arbitrary choice primarily dependent on the form/type of PBS used to convert the polarization encoded qubits to which path encoded qubits. Any other choice is also acceptable and it does not have to be symmetric between the two qubits. Our solution in Eqn (10) depends on the value of the measured quadrature $`X`$. Simple local rotations using phase shifters dependent on the measurement result $`X`$ can be performed via a feed forward process to transform this state to $`\beta _1|H_a|V_b+\beta _2|V_a|H_b`$ which is independent of $`X`$. This does mean our homodyne measurement must be accurate enough such that we can determine $`\varphi (X)`$ precisely, otherwise this unwanted phase factor cannot be undone. By this we mean that the uncertainty in the $`X`$ quadrature homodyne measurement must be much less than $`2\pi /\alpha _c\mathrm{sin}\theta `$ and this can generally be achieved by ensuring that the strength of the local oscillator is much more intense than the probe mode. In the above solutions (9) and (10) we have used the the approximate symbol $``$ as there is a small but finite probability that the state (9) can occur for $`X<X_0`$ and vice versa. The probability of this error occurring is given by $`P_{\mathrm{error}}={\displaystyle \frac{1}{2}}\mathrm{Erfc}[X_d/2\sqrt{2}]`$ (11) which is less than $`10^4`$ when the distance $`X_d\alpha _c\theta ^2>8`$. This shows that it is possible to operate in the regime of weak cross-Kerr non-linearities ($`\theta \pi `$) and achieve an effectively deterministic parity measurement. ## 4 Bell state measurements These effectively deterministic nondestructive parity measurements are critically important in optical quantum information processing as they naturally allow an efficient and deterministic Bell state measurement to be implemented. Bell state measurements are known to be one of the tools and mechanism required in quantum computation and communication. The four Bell states can be written as $`|\mathrm{\Psi }^\pm `$ $``$ $`{\displaystyle \frac{1}{\sqrt{2}}}(|H,V\pm |V,H)|\mathrm{\Phi }^\pm |{\displaystyle \frac{1}{\sqrt{2}}}(|H,H\pm |V,V)`$ (12) and we can now see why the parity gate can form the basis of a Bell state detector. The parity gate distinguishes states within the even parity $`|H,H`$ and $`|V,V`$ subspace from the odd parity $`|H,V`$ and $`|V,H`$ subspace. Hence one application of the parity detector distinguishes two of the Bell states $`|\mathrm{\Phi }^\pm `$ from the $`|\mathrm{\Psi }^\pm `$ ones without destroying them. Similarly if we replace the polarizing beam-splitter in the parity gates with 45 degree PBS then the parity gate will allow us to distinguish the $`|\mathrm{\Phi }^+,|\mathrm{\Psi }^+`$ Bell states from $`|\mathrm{\Phi }^{},|\mathrm{\Psi }^{}`$ ones. Since both of these detectors are nondestructive on the qubits and select different pairs of Bell states, they allow the natural construction of a Bell state detector (depicted in Fig (4)). From each parity measurements we get one of information indicating whether the parity was even or odd and so from both parity measurements we end up with four possible results (even, even), (even, odd), (odd, even) and (odd, odd). This is enough to uniquely identify all the Bell states as the $`|\mathrm{\Phi }^+`$ gives the result (even, even), $`|\mathrm{\Phi }^{}`$ (even, odd), $`|\mathrm{\Psi }^+`$ (odd, even) and $`|\mathrm{\Psi }^{}`$ (odd, odd). It is important that after an odd parity measurement result that we remove the unwanted phase factors that have arisen. This need to be done in the same basis as the PBS in the particular parity gate. For instance for an odd parity results giving $`X=X_1`$ on the first parity gate a phase shift $`\varphi (X_1)`$ needs to be removed in the PBS $`\{H,V\}`$ basis. Similarly for a odd parity results giving $`X=X_2`$ on the second parity gate a phase shift $`\varphi (X_2)`$ needs to be removed in the PBS $`\{H+V,HV\}`$ basis So far we have shown how it is possible using linear elements, weak cross-Kerr non-linearities and homodyne measurements to create a wide range of high efficiency quantum detectors and gates that can perform task ranging from photon number discrimination to Bell state measurements. This is all achieved non-destructively on the photonic qubits and hence provides a critical set of tools extremely useful for single photon quantum computation and communication. With these tools universal quantum computation can be achieved using the ideas and techniques originally proposed by KLM. ## 5 A resource efficient CNOT gate The parity gate and Bell state detector has shown how versatile the weak non-linearities and homodyne conditioning measurements are. Both of these gates/detectors can be used to induce two qubit operations and hence are all they are necessary with single qubit operation and single photon measurements to perform universal quantum computation. The parity and Bell state gates are not the typical two qubits gates that one generally considers in the standard quantum computational models. The typical two qubit gate generally considered is the CNOT gate. This can be constructed from two parity gates (like the Bell state detector) but it also requires an ancilla qubit. This CNOT gate is depicted schematically shown in Fig (5)) and operates as follows. Assume that our control and target qubits are initially prepared as $`c_0|H_c+c_1|V_c`$ and $`d_0|H_t+d_1|V_t`$. With an ancilla qubit prepared as $`|H_a+|V_a`$ the action of the first parity gate on the control and ancilla qubits (with appropriate phase corrections for the odd parity result) conditions the system to $`\left[c_0|HH_{ca}+c_1|VV_{ca}\right]\left[d_0|H_t+d_1|V_t\right]`$ (13) The action of the second parity gate (using 45 deg PBS’s instead of normal PBS’s) on the ancilla qubit and target qubit conditions the three qubit system to $`\left\{c_0|H_cc_1|V_c\right\}(d_0d_1)|\overline{D},\overline{D}_{at}+\left\{c_0|H_c+c_1|V_c\right\}(d_0+d_1)|D,D_{at}`$ (14) where $`|D=|H+|V`$, $`|\overline{D}=|H|V`$ and for the odd parity measurement result $`X<X_0`$ the usual phase correction is applied. Also for this odd parity result a bit flip is applied to the ancilla qubit and a sign flip $`|V_c|V_c`$ on the control qubit. Once this operation have been performed the ancilla mode is measured in the $`\{H,V\}`$ basis using QND photon number resolving detectors. The output state of the control and target qubits is then final state from these interactions and feed forward $`c_0d_0|HH_{ct}+c_0d_1|HV_{ct}+c_1d_0|VV_{ct}+c_1d_1|VH_{ct},`$ (15) where an additional bit flip was applied to the target qubit if the ancilla photons state was $`|V`$. This final state is the state that one will expect after a CNOT gate is applied to the initial control and target qubits. This really shows that our QND-based parity gates have performed a near deterministic CNOT operation utilizing only one ancilla qubit (which is not destroyed at the end of the gate). This represents a huge saving in the physical resources to implement single photon quantum logic gates. ## 6 Concluding Discussions We have shown how it is possible to create near deterministic two qubit gates (parity, bell and CNOT) without a huge overhead in ancilla resources. In fact, an ancilla photon is required only for the CNOT gate. The key addition to the general linear optical resources are weak cross-Kerr nonlinearities and efficient homodyne measurements. Homodyne measurements are a well established technique frequently used in the continuous variable quantum information processing community. However weak cross-Kerr nonlinearities are not commonly used elements within optical quantum computational devices and as such it a discussion of the source and strength of such elements is required. We will start with a discussion of the strength of the nonlinearity as this constraints the possible physical realisations, however before this we really need to define what we mean by weak or weak compared with what. Basically it is well known that deterministic two qubit gates can be performed if one has access to a cross-Kerr nonlinearity that can induce a $`\pi `$ phase shift directly between single photon. This leads to a natural definition of weak nonlinearities, that is, the use of nonlinear cross-Kerr materials (when all are taken into account) that can not directly induce a phase shift within an order of magnitude or several orders of magnitude of $`\pi `$. This seems to give an acceptable functional definition. For the parity based gates discussed previously we have established that the nonlinearity $`\theta `$ must satisfy the constraint $`\alpha _c\theta ^28`$ where just to re-emphasise $`\alpha _c`$ is the amplitude of the probe beam. Thus due to the weak nature of the nonlinearity $`\theta 1`$ we must choose $`\alpha _c10/\theta ^2`$, so for instance if $`\theta 10^2`$ then $`\alpha _c10^5`$ (which corresponds to a probe beam with mean photon number $`10^{10}`$). For a smaller $`\theta `$ we need a much larger $`\alpha _c`$. This puts a natural constraints on $`\theta `$, since $`\alpha _c`$ can not be made arbitrarily large in practice. This leads to the question of a mechanism to achieve the weak cross-Kerr nonlinearity. Natural $`\chi ^3`$ materials have small nonlinearities on the order of $`10^{18}`$ which would require lasers with $`\alpha _c10^{37}`$ which is physically unrealistic. However, systems such as optical fibers , silica whispering-gallery microresonators and cavity QED systems , and EIT are capable of producing much much larger nonlinearities. For instance calculations for EIT systems in NV diamond have shown potential phase shifts of order of magnitude of $`\theta =0.01`$. With $`\theta =0.01`$ the probe beam must have an amplitude of at least $`10^5`$ which is physically reasonable with current technology. Finally, by using these weak cross-Kerr nonlinearities to aid in the construction of near deterministic two qubit gates we can build quantum circuits with far fewer resources than is known for the current corresponding linear optical only approaches.It is straightforward to show in principle that an $`n`$ qubit computation requires only of order $`n`$ single photons sources. This has enormous implications for the development of single photon quantum computing and information processing devices and truly indicates the power of a little nonlinearity. Acknowledgments: This work was supported by the European Project RAMBOQ. KN acknowledges support in part from MPHPT and JSPS. WJM acknowledges support in the form of a JSPS fellowship. ## 7 References
warning/0506/math0506198.html
ar5iv
text
# More on Reverse Triangle Inequality in Inner Product Spaces11footnote 12000 Mathematics Subject Classification. Primary 46C05; Secondary 26D15. Key words and phrases. Triangle inequality, reverse inequality, Schwarz inequality, inner product space. ## 1 Introduction. It is interesting to know under which conditions the triangle inequality went the other way in a normed space $`X`$; in other words, we would like to know if there is a constant $`c`$ with the property that $`c_{k=1}^nx_k_{k=1}^nx_k`$ for any finite set $`x_1,\mathrm{},x_nX`$. M. Nakai and T. Tada proved that the normed spaces with this property are precisely those of finite dimensional. The first authors investigating reverse of the triangle inequality in inner product spaces were J. B. Diaz and F. T. Metcalf by establishing the following result as an extension of an inequality given by M. Petrovich for complex numbers: Diaz-Metcalf Theorem. Let $`a`$ be a unit vector in an inner product space $`(H;.,.)`$. Suppose the vectors $`x_kH,k\{1,\mathrm{},n\}`$ satisfy $`0r{\displaystyle \frac{Rex_k,a}{x_k}},k\{1,\mathrm{},n\}`$ Then $`r{\displaystyle \underset{k=1}{\overset{n}{}}}x_k{\displaystyle \underset{k=1}{\overset{n}{}}}x_k.`$ where equality holds if and only if $`{\displaystyle \underset{k=1}{\overset{n}{}}}x_k=r{\displaystyle \underset{k=1}{\overset{n}{}}}x_ka.`$ Inequalities related to the triangle inequality are of special interest; cf. Chapter XVII of . They may be applied to get interesting inequalities in complex numbers or to study vector-valued integral inequalities $`\text{[3]},\text{[4]}`$. Using several ideas and following the terminology of and we modify or refine some results of Dragomir and ours and get some new reverses of triangle inequality. Among several results, we show that if $`a`$ is a unit vector in a real or complex inner product space $`(H;.,.)`$, $`x_kH\{0\},1kn`$, $`\alpha =\mathrm{min}\{x_k:1kn\},p(0,\sqrt{\alpha ^2+1}),\mathrm{max}\{x_ka:1kn\}p`$ and $`\beta =\mathrm{min}\{\frac{x_k^2p^2+1}{2x_k}:1kn\}`$, then $$\underset{k=1}{\overset{n}{}}x_k\underset{k=1}{\overset{n}{}}x_k\frac{1\beta }{\beta }Re\underset{k=1}{\overset{n}{}}x_k,a.$$ We also examin some reverses for the celebrated Schwarz inequality. In particular, it is proved that if $`a`$ is a unit vector in a real or complex inner product space $`(H;.,.)`$, $`r,s>0,p(0,s],D=\{xH,rxsap\},x_1,x_2D\{0\}`$ and $`\alpha _{r,s}=\mathrm{min}\{\frac{r^2x_k^2p^2+s^2}{2rsx_k}:1k2\}`$, then $$\frac{x_1x_2Rex_1,x_2}{(x_1+x_2)^2}\alpha _{r,s}$$ Throughout the paper $`(H;.,.)`$ denotes a real or complex inner product space. We use repeatedly the Cauchy-Schwarz inequality without mentioning it. The reader is referred to $`\text{[9]},\text{[5]}`$ for the terminology on inner product spaces. ## 2 Reverse of Triangle inequality. We start this section by pointing out the following theorem of which is a modification of theorem 3 of . ###### Theorem 2.1 Let $`a_1,\mathrm{},a_m`$ be orthonormal vectors in the complex inner product space $`(H;.,.)`$. Suppose that for $`1tm,r_t,\rho _tR`$ and that the vectors $`x_kH,k\{1,\mathrm{},n\}`$ satisfy $$0r_t^2x_kRex_k,r_ta_t,0\rho _t^2x_kImx_k,\rho _ta_t,1tm$$ (1) Then $$(\underset{t=1}{\overset{m}{}}(r_t^2+\rho _t^2))^{\frac{1}{2}}\underset{k=1}{\overset{n}{}}x_k\underset{k=1}{\overset{n}{}}x_k$$ (2) and the equality holds in $`(2)`$ if and only if $$\underset{k=1}{\overset{n}{}}x_k=\underset{k=1}{\overset{n}{}}x_k\underset{t=1}{\overset{m}{}}(r_t+i\rho _t)a_t.$$ (3) The following theorem is a strengthen of Corollary 1 of and a generalization of Theorem 2 of . ###### Theorem 2.2 Let $`a`$ be a unit vector in the complex inner product space $`(H;.,.)`$. Suppose that the vectors $`x_kH\{0\},k\{1,\mathrm{},n\}`$ satisfy $$\mathrm{max}\{rx_ksa:1kn\}p,\mathrm{max}\{r^{}x_kis^{}a:1kn\}q$$ where $`r,r^{},s,s^{}>0`$ and $$p((r\alpha )^2+s^2)^{\frac{1}{2}},q((r^{}\alpha )^2+s^2)^{\frac{1}{2}}$$ and $$\alpha =\mathrm{min}\{x_k:1kn\}.$$ Let $$\alpha _{r,s}=\mathrm{min}\{\frac{r^2x_k^2p^2+s^2}{2rsx_k}:1kn\},$$ $$\beta _{r^{},s^{}}=\mathrm{min}\{\frac{r^2x_k^2q^2+s^2}{2r^{}s^{}x_k}:1kn\}$$ Then $$(\alpha _{r,s}^2+\beta _{r^{},s^{}}^2)^{\frac{1}{2}}\underset{k=1}{\overset{n}{}}x_k\underset{k=1}{\overset{n}{}}x_k$$ and the equality holds if and only if $$\underset{k=1}{\overset{n}{}}x_k=(\alpha _{r,s}+i\beta _{r^{},s^{}})\underset{k=1}{\overset{n}{}}x_ka.$$ Proof. From the first inequality above we infer that $$rx_ksa,rx_ksap^2$$ $$r^2x_k^2+s^2p^22Rerx_k,sa.$$ Then $$\frac{r^2x_k^2+s^2p^2}{2rsx_k}x_kRex_k,a.$$ Similarly $$\frac{r^2x_k^2q^2+s^2}{2r^{}s^{}x_k}x_kImx_k,a$$ consequently $$\alpha _{r,s}x_kRex_k,a$$ and $$\beta _{r^{},s^{}}x_kImx_k,a.$$ Applying Theorem 2.1 for $`m=1,r_1=\alpha _{r,s}`$ and $`\rho _1=\beta _{r^{},s^{}}`$ we deduce desired inequality. $`\mathrm{}`$ The next result is an extension of Corollary 3 of . ###### Corollary 2.3 Let $`a`$ be a unit vector in the complex inner product space $`(H;.,.)`$. Suppose that $`x_kH,k\{1,\mathrm{},n\},\mathrm{max}\{rx_ksa:1kn\}r,\mathrm{max}\{rx_kisa:1kn\}s`$ where $`r>0,s>0`$ and $`\alpha =\mathrm{min}\{x_k:1kn\}.`$ Then $$\frac{r\alpha }{s\sqrt{2}}\underset{k=1}{\overset{n}{}}x_k\underset{k=1}{\overset{n}{}}x_k.$$ The equality holds if and only if $$\underset{k=1}{\overset{n}{}}x_k=r\alpha \frac{(1+i)}{2s}\underset{k=1}{\overset{n}{}}x_ka.$$ Proof. Apply Theorem 2.2 with $`r=r^{},s=s^{},p=r,q=s`$. Note that $`\alpha _{r,s}=\frac{r\alpha }{2s}=\beta _{r^{},s^{}}.\mathrm{}`$ ###### Theorem 2.4 Let $`a`$ be a unit vector in $`H;.,.)`$. Suppose that the vectors $`x_kH,k\{1,\mathrm{},n\}`$ satisfy $$\mathrm{max}\{rx_ksa:1kn\}p<((r\alpha )^2+s^2)^{\frac{1}{2}}$$ where $`r>0,s>0`$ and $$\alpha =\underset{1kn}{\mathrm{min}}x_k.$$ Let $$\alpha _{r,s}=\mathrm{min}\{\frac{r^2x_k^2p^2+s^2}{2rsx_k}:1kn\}.$$ Then $$\alpha _{r,s}\underset{k=1}{\overset{n}{}}x_k\underset{k=1}{\overset{n}{}}x_k.$$ Moreover, the equality holds if and only if $$\underset{k=1}{\overset{n}{}}x_k=\alpha _{r,s}\underset{k=1}{\overset{n}{}}x_ka.$$ Proof. Proof is similar to that of Theorem 2.2 in which we use Theorem 2.1 with $`m=1,\rho _1=0.\mathrm{}`$ ###### Theorem 2.5 Let $`a`$ be a unit vector in $`(H;.,.)`$. Suppose that $`r,s>0`$, and vectors $`x_kH\{0\},k\{1,\mathrm{},n\}`$ satisfy $$\underset{k=1}{\overset{n}{}}x_k=0.$$ Then $$\sqrt{r^2\alpha ^2+s^2}\mathrm{max}\{rx_ksa:1n\}$$ where $$\alpha =\underset{1kn}{\mathrm{min}}x_k.$$ Proof. Let $`p=\mathrm{max}\{rx_ksa:1kn\}`$. If $`p<\sqrt{r^2\alpha ^2+s^2}`$, then using Theorem 2.4 we get $$\alpha _{r,s}\underset{k=1}{\overset{n}{}}x_k\underset{k=1}{\overset{n}{}}x_k=0.$$ Hence $`\alpha _{r,s}=0`$. On the other hand $`\frac{p^2s^2}{r^2}<\alpha ^2`$, so $$\alpha _{r,s}=\mathrm{min}\{\frac{r^2x_k^2p^2+s^2}{2rsx_k}:1kn\}>0$$ a contradiction. $`\mathrm{}`$ ###### Theorem 2.6 Let $`a_1,\mathrm{},a_m`$ be orthonormal vectors in the complex inner product space $`(H;.,.),M_tm_t>0,L_t\mathrm{}_t>0,1tm`$ and $`x_kH\{0\},k\{1,\mathrm{},n\}`$ such that $$ReM_ta_tx_k,x_km_ta_t0,ReL_tia_tx_k,x_k\mathrm{}_tia_t0$$ or equivalently $$x_k\frac{m_t+M_t}{2}a_t\frac{M_tm_t}{2},x_k\frac{L_t+\mathrm{}_t}{2}ia_t\frac{L_t\mathrm{}_t}{2}$$ for all $`1kn`$ and $`1tm`$. Let $$\alpha _{m_t,M_t}=\mathrm{min}\{\frac{x_k^2+m_tM_t}{(m_t+M_t)x_k}:1kn\},1tm$$ and $$\alpha _{\mathrm{}_t,L_t}=\mathrm{min}\{\frac{x_k^2+\mathrm{}_tL_t}{(m_t+M_t)x_k}:1kn\},1tm$$ then $$(\underset{t=1}{\overset{m}{}}\alpha _{m_t,M_t}^2+\alpha _{\mathrm{}_t,L_t}^2)^{\frac{1}{2}}\underset{k=1}{\overset{n}{}}x_k\underset{k=1}{\overset{n}{}}x_k.$$ The equality holds if and only if $$\underset{k=1}{\overset{n}{}}x_k=(\underset{k=1}{\overset{n}{}}x_k)\underset{t=1}{\overset{m}{}}(\alpha _{m_t,M_t}+i\alpha _{\mathrm{}_t,L_t})a_t.$$ Proof. Given $`1tm`$ and all $`1kn`$, it follows from $`x_k\frac{m_t+M_t}{2}a_t\frac{M_tm_t}{2}`$ that $$x_k^2+m_tM_t(m_t+M_t)Rex_k,a_t.$$ Then $$\frac{x_k^2+m_tM_t}{(m_t+M_t)x_k}x_kRex_k,a_t$$ and so $$\alpha _{m_t,M_t}x_kRex_k,a_t.$$ Similarly from the second inequality we deduce that $$\alpha _{\mathrm{}_t,L_t}x_kImx_k,a_t.$$ Applying Theorem 2.5 for $`r_t=\alpha _{m_t,M_t}`$ and $`\rho _t=\alpha _{\mathrm{}_t,L_t}`$, we obtain the required inequality. $`\mathrm{}`$ We will need theorem 7 of . We mention it for the sake of completeness. ###### Theorem 2.7 Let $`a`$ be a unit vector in $`(H;.,.)`$, and $`x_kH\{0\},k\{1,\mathrm{},n\}`$. If $`r_k0,k\{1,\mathrm{},n\}`$ such that $$x_kRex_k,ar_k$$ then $$\underset{k=1}{\overset{n}{}}x_k\underset{k=1}{\overset{n}{}}x_k\underset{k=1}{\overset{n}{}}r_k.$$ The equality holds if and only if $$\underset{k=1}{\overset{n}{}}x_k\underset{k=1}{\overset{n}{}}r_k$$ and $$\underset{k=1}{\overset{n}{}}x_k=(\underset{k=1}{\overset{n}{}}x_k\underset{k=1}{\overset{n}{}}r_k)a.$$ ###### Theorem 2.8 Let $`a`$ be a unit vector in $`(H;.,.)`$ and $`x_kH\{0\},k\{1,\mathrm{},n\}.`$ Let $$\alpha =\mathrm{min}\{x_k:1kn\},p(0,\sqrt{\alpha ^2+1}),\mathrm{max}\{x_ka:1kn\}p$$ and $$\beta =\mathrm{min}\{\frac{x_k^2p^2+1}{2x_k}:1kn\}.$$ Then $$\underset{k=1}{\overset{n}{}}x_k\underset{k=1}{\overset{n}{}}x_k\frac{1\beta }{\beta }Re\underset{k=1}{\overset{n}{}}x_k,a.$$ The equality holds if and only if $$\underset{k=1}{\overset{n}{}}x_k\frac{1\beta }{\beta }Re\underset{k=1}{\overset{n}{}}x_k,a$$ and $$\underset{k=1}{\overset{n}{}}x_k=(\underset{k=1}{\overset{n}{}}x_k\frac{1\beta }{\beta }Re\underset{k=1}{\overset{n}{}}x_k,a)a.$$ Proof. Since $`\mathrm{max}\{x_ka:1kn\}p`$, we have $$x_ka,x_kap^2$$ $$x_k^2+1p^22Rex_k,a$$ $$\frac{x_k^2p^2+1}{2x_k}x_kRex_k,a$$ $$\beta x_kRex_k,a$$ $$x_k\frac{1}{\beta }Rex_k,a$$ for all $`k\{1,\mathrm{},n\}`$. Then $$x_kRex_k,a\frac{1\beta }{\beta }Rex_k,a,k\{1,\mathrm{},n\}.$$ Applying Theorem 2.7 for $`r_k=\frac{1\beta }{\beta }Rex_k,a,k\{1,\mathrm{},n\}`$, we deduce the desired inequality. $`\mathrm{}`$ As a corollary, we obtain a result similar to Theorem 9 of : ###### Corollary 2.9 Let $`a`$ be a unit vector in $`(H;.,.)`$ and $`x_kH\{0\},k\{1,\mathrm{},n\}`$. Let $$\mathrm{max}\{x_ka:1kn\}1$$ and $$\alpha =\mathrm{min}\{x_k:1kn\}.$$ Then $$\underset{k=1}{\overset{n}{}}x_k\underset{k=1}{\overset{n}{}}x_k\frac{2\alpha }{\alpha }Re\underset{k=1}{\overset{n}{}}x_k,a.$$ The equality holds if and only if $$\underset{k=1}{\overset{n}{}}x_k\frac{2\alpha }{\alpha }Re\underset{k=1}{\overset{n}{}}x_k,a$$ and $$\underset{k=1}{\overset{n}{}}x_k=(\underset{k=1}{\overset{n}{}}x_k\frac{2\alpha }{\alpha }Re\underset{k=1}{\overset{n}{}}x_k,a)a.$$ Proof. Apply Theorem 2.8 with $`\beta =\frac{\alpha }{2}.\mathrm{}`$ ###### Theorem 2.10 Let $`a`$ be a unit vector in $`(H;.,.)`$, $`Mm>0`$ and $`x_kH\{0\},k\{1,\mathrm{},n\}`$ such that $$ReMax_k,x_kma0$$ or equivalently $$x_k\frac{m+M}{2}a\frac{Mm}{2}.$$ Let $$\alpha _{m,M}=\mathrm{min}\{\frac{x_k^2+mM}{(m+M)x_k}:1kn\}.$$ Then $$\underset{k=1}{\overset{n}{}}x_k\underset{k=1}{\overset{n}{}}x_k\frac{1\alpha _{m,M}}{\alpha _{m,M}}Re\underset{k=1}{\overset{n}{}}x_k,a.$$ The equality holds if and only if $$\underset{k=1}{\overset{n}{}}x_k\frac{1\alpha _{m,M}}{\alpha _{m,M}}Re\underset{k=1}{\overset{n}{}}x_k,a$$ and $$\underset{k=1}{\overset{n}{}}x_k=(\underset{k=1}{\overset{n}{}}x_k\frac{1\alpha _{m,M}}{\alpha _{m,M}}Re\underset{k=1}{\overset{n}{}}x_k,a)a.$$ Proof. For each $`1kn`$, it follows from the inequality $$x_k\frac{m+M}{2}a\frac{Mm}{2}$$ that $$x_k\frac{m+M}{2}a,x_k\frac{m+M}{2}a(\frac{Mm}{2})^2.$$ Hence $$x_k^2+mM(m+M)Rex_k,a.$$ So that $$\alpha _{m,M}x_kRex_k,a$$ consequently $$x_kRex_k,a\frac{1\alpha _{m,M}}{\alpha _{m,M}}Rex_k,a.$$ Now apply Theorem 2.7 for $`r_k=\frac{1\alpha _{m,M}}{\alpha _{m,M}}Rex_k,a,k\{1,\mathrm{},n\}.\mathrm{}`$ ## 3 Reverses of Schwarz inequality. In this section we provide some reverses of the Schwarz inequality. The first theorem is an extension of Proposition 5.1 of . ###### Theorem 3.1 Let $`a`$ be a unit vector in $`(H;.,.)`$. Suppose that $`r,s>0,p(0,s]`$ and $$D=\{xH,rxsap\}.$$ If $`0x_1D,0x_2D`$, then $$\frac{x_1x_2Rex_1,x_2}{(x_1+x_2)^2}\frac{1}{2}(1(\frac{r^2x_1^2p^2+s^2}{2rsx_1})^2)$$ or $$\frac{x_1x_2Rex_1,x_2}{(x_1+x_2)^2}\frac{1}{2}(1(\frac{r^2x_2^2p^2+s^2}{2rsx_2})^2)$$ Proof. Put $`\alpha _{r,s}=\mathrm{min}\{\frac{r^2x_k^2p^2+s^2}{2rsx_k}:1k2\}`$. By Theorem 2.4, we obtain $$\alpha _{r,s}(x_1+x_2)x_1+x_2.$$ Then $$\alpha _{r,s}^2(x_1^2+2x_1x_2+x_2^2)x_1^2+2Rex_1,x_2+x_2^2.$$ Set $`\alpha _{r,s}^2=1t^2`$. Then $$\frac{x_1x_2Rex_1,x_2}{(x_1+x_2)^2}\frac{1}{2}t^2,$$ namely $$\frac{x_1x_2Rex_1,x_2}{(x_1+x_2)^2}\frac{1}{2}(1\alpha _{r,s}^2).\mathrm{}$$ ###### Corollary 3.2 Let $`a`$ be a unit vector in $`(H;.,.)`$. Suppose that $`r,s>0`$ and $$D=\{xH,rxsas\}.$$ If $`x,yD`$ and $`0<x<y`$, then $$\frac{xyRex,y}{(x+y)^2}\frac{1}{2}(1(\frac{rx}{2s})^2).$$ Proof. In the notation of the proof of Theorem 3.1 we get from $`p=s,x_1=x,x_2=y`$ that $`\alpha _{r,s}=\frac{rx}{2s}`$. Now apply Theorem 3.1. $`\mathrm{}`$ ###### Corollary 3.3 Let $`a`$ be a unit vector in $`(H;.,.)`$, $`Mm>0`$ and $`x_kH\{0\},k=1,2`$ such that $$ReMax_k,x_kma0$$ or equivalently, $$x_k\frac{m+M}{2}a\frac{Mm}{2}.$$ Then $$\frac{x_1x_2Rex_1,x_2}{(x_1+x_2)^2}\frac{1}{2}(1(\frac{x_1^2+mM}{(m+M)x_1})^2)$$ or $$\frac{x_1x_2Rex_1,x_2}{(x_1+x_2)^2}\frac{1}{2}(1(\frac{x_2^2+mM}{(m+M)x_2})^2).$$ Proof. Put $`r=1,s=\frac{m+M}{2},p=\frac{Mm}{2},x=x_1`$ and $`y=x_2`$ in Theorem 3.1. $`\mathrm{}`$ Acknowledgement. The authors would like to thank the referees for their valuable suggestions. Arsalan Hojjat Ansari Mohammad Sal Moslehian Dept. of Math., Ferdowsi Univ., P. O. Box 1159, Mashhad 91775, Iran E-mail: moslehian@ferdowsi.um.ac.ir Home: http://www.um.ac.ir/$``$moslehian/
warning/0506/hep-ph0506118.html
ar5iv
text
# Properties of the Effective Hamiltonian for the System of Neutral Kaons11footnote 1 Paper presented at The XXXVI Symposium on Mathematical Physics, Poster session, Toruń, Poland, June 9-12, 2004 . This is shortened version of [1] ## 1 Introduction Following the LOY approach, a nonhermitian Hamiltonian $`H_{}`$ is usually used to study the properties of the particle-antiparticle unstable system \- $$H_{}M\frac{i}{2}\mathrm{\Gamma },$$ (1) where $$M=M^+,\mathrm{\Gamma }=\mathrm{\Gamma }^+$$ (2) are $`(2\times 2)`$ matrices acting in a two-dimensional subspace $`_{}`$ of the total state space $``$. The $`M`$-matrix is called the mass matrix and $`\mathrm{\Gamma }`$ is the decay matrix. Lee, Oehme and Yang derived their approximate effective Hamiltonian $`H_{}H_{LOY}`$ by adapting the one-dimensional Weisskopf-Wigner (WW) method to the two-dimensional case corresponding to the neutral kaon system. Almost all properties of this system can be described by solving the Schrödinger-like equation \- $$i\frac{}{t}|\psi ;t_{}=H_{}|\psi ;t_{},(tt_0>\mathrm{})$$ (3) (where we have used $`\mathrm{}=c=1`$) with the initial conditions $$|\psi ;t=t_0_{}=1,|\psi ;t_0=0_{}=0,$$ (4) for $`|\psi ;t=t_0_{}`$ belonging to the subspace of states $`_{}`$ ($`_{}`$), spanned by, e.g., orthonormal neutral kaons states $`K^0`$ and $`\overline{K}^0`$. The solutions of Eq. (3) may be written in a matrix form, which may be used to define the time evolution operator $`U_{}(t)`$ acting in subspace $`_{}`$ $$|\psi ;t_{}=U_{}(t)|\psi ;t_0=0_{}U_{}(t)|\psi _{},$$ (5) where $$|\psi _{}a_1|\text{1}+a_2|\text{2}$$ (6) and $`|\text{1}`$ denotes particle "1" – in the present case $`|K^0`$ whereas $`|\text{2}`$ corresponds to the antiparticle state for particle ”1”: $`|\overline{K}^0,`$ $`\text{j}|\text{k}=\delta _{jk},`$ $`j,k=1,2`$. It is usually assumed that the real parts of the diagonal matrix elements of $`H_{}`$, namely $`\mathrm{}(),`$ $$\mathrm{}(h_{jj})M_{jj}(j=1,2),$$ (7) where $$h_{jk}=\text{j}|H_{}|\text{k}(j,k=1,2)$$ (8) correspond to the masses of the particle "1" and its antiparticle "2" \- . $`\mathrm{}()`$ is the imaginary part of $`h_{jj}`$ $$\mathrm{}(h_{jj})\mathrm{\Gamma }_{jj}(j=1,2)$$ (9) and $`\mathrm{\Gamma }_{jj}`$ are interpreted as the decay widths of the particles. According to the standard result of the LOY approach, in a CPT invariant system, i.e. when $$\mathrm{\Theta }H\mathrm{\Theta }^1=H,$$ (10) (where $`\mathrm{\Theta }=CPT`$, $`H=H^+`$ is the Hamiltonian of the total system under consideration) we have $$h_{11}^{LOY}=h_{22}^{LOY}.$$ (11) The universal properties of the unstable particle-antiparticle subsystem described by the $`H`$ fulfilling the condition (10), may be investigated by using the matrix elements of the exact $`U_{},`$ instead of the approximate one used in the LOY theory. The exact $`U_{}`$ can be written as follows $$U_{}(t)=PU(t)P,$$ (12) where $$P|\text{1}\text{1}|+|\text{2}\text{2}|,$$ (13) and $`U(t)`$ is the exact evolution operator acting in the whole state space. This operator is the solution of the Schrödinger equation $$i\frac{}{t}U(t)|\varphi =HU(t)|\varphi ,U(0)=I.$$ (14) $`I`$ is the unit operator in the $``$ space and $`|\varphi |\varphi ;t_0=0`$ is the initial state of the system. In the remaining part of the poster we will be using the following matrix representation of the evolution operator $$U_{}(t)\left(\begin{array}{cc}\text{A(t)}& \text{0}\\ \text{0}& \text{0}\end{array}\right),$$ (15) where 0 denotes the zero submatrices of the suitable dimension, and the A(t) is a $`(2\times 2)`$ matrix acting in $`_{}`$ $$\text{A(t)}=\left(\begin{array}{cc}A_{11}(t)\hfill & A_{12}(t)\hfill \\ A_{21}(t)\hfill & A_{22}(t)\hfill \end{array}\right),$$ (16) where $$A_{jk}(t)=\text{j}|U_{}(t)|\text{k}\text{j}|U(t)|\text{k}(j,k=1,2).$$ (17) Assuming that the property (10) holds and using the following definitions $$\mathrm{\Theta }|\text{1}e^{i\theta }|\text{2},\mathrm{\Theta }|\text{2}e^{i\theta }|\text{1},$$ (18) it can be shown that $$A_{11}(t)=A_{22}(t).$$ (19) A very important relation between the amplitudes $`A_{12}(t)`$ and $`A_{21}(t)`$ follows from the famous Khalfin Theorem \- $$r(t)\frac{A_{12}(t)}{A_{21}(t)}=constr|r|=1.$$ (20) General conclusions concerning the properties of the matrix elements of $`H_{}`$ can be drawn by analyzing the following identity $$H_{}(t)i\frac{\text{A}(t)}{t}[\text{A}(t)]^1.$$ (21) Using Eq. (21) we can easily find the general formulae for the diagonal matrix elements $`h_{jj}`$, of $`H_{}(t)`$ and next assuming (10) and using relation (19) which follows from our earlier assumptions, we get $$h_{11}(t)h_{22}(t)=\frac{i}{det\text{A}(t)}(\frac{A_{21}(t)}{t}A_{12}(t)\frac{A_{12}(t)}{t}A_{21}(t)).$$ (22) In it was shown, by using relation (22), that this result means that in the considered case (with CPT conserved) for $`t>0`$ we get the following theorem $$h_{11}(t)h_{22}(t)=0\frac{A_{12}(t)}{A_{21}(t)}=const(t>0).$$ (23) Thus, for $`t>0`$ the problem under study is reduced to the Khalfin Theorem (see relation (20)) . Having noticed this, let us now turn our attention to the conclusions following from Khalfin’s Theorem. $`CP`$ noninvariance requires that $`|r|1`$ . This means that in this case the following condition must be fulfilled: $`r=r(t)const.`$ Consequently, if in the considered system property (10) holds, but at the same time $$[𝒞𝒫,H]0$$ (24) and the unstable states "1" i "2" are connected by (18), then in this system for $`t>0`$ $$h_{11}(t)h_{22}(t)0.$$ (25) So, in the exact quantum theory the difference $`(h_{11}(t)h_{22}(t))`$ cannot be equal to zero with CPT conserved and CP violated. ## 2 A model: one pole approximation While describing the two and three pion decay we are mostly interested in the $`|K_S`$ and $`|K_L`$ superposition of $`|K^0`$ and $`|\overline{K}^0.`$ These states correspond to the physical $`|K_S`$ and $`|K_L`$ neutral kaon states $`|K_S=p|K^0+q|\overline{K}^0,|K_L=p|K^0q|\overline{K}^0.`$ (26) Using the spectral formalism we can write an unstable state $`|\lambda (t)`$ as $`|\lambda (t)={\displaystyle \underset{q}{}}|q(t)\omega _\lambda (q),`$ (27) where $`|q(t)=e^{itH}|q`$, vectors $`|q`$ form a complete set of eigenvectors of the hermitian, quantum-mechanical Hamiltonian $`H`$ and $`\omega _\lambda (q)=q|\lambda .`$ If the continuous eigenvalue is denoted by $`m`$, we can define the survival amplitude $`A(t)`$ (or the transition amplitude in the case of $`K^0\overline{K}^0`$ ) in the following way: $`A(t)={\displaystyle \underset{Spec(H)}{}}𝑑me^{imt}\rho (m),`$ (28) where the integral extends over the whole spectrum of the Hamiltonian and density $`\rho (m)`$ is defined as follows $`\rho (m)=|\omega _\lambda (m)|^2,`$ (29) where $`\omega _\lambda (m)=m|\lambda .`$ In accordance with formula (27) the unstable states $`K_S`$ and $`K_L`$ may now be written as a superposition of the eigenkets $`|K_S={\displaystyle _0^{\mathrm{}}}𝑑m{\displaystyle \underset{\alpha }{}}\omega _{S,\alpha }(m)|\varphi _\alpha (m);`$ (30) $`|K_L={\displaystyle _0^{\mathrm{}}}𝑑m{\displaystyle \underset{\beta }{}}\omega _{L,\beta }(m)|\varphi _\beta (m).`$ (31) The Breit-Wigner ansatz $`\rho _{WB}(m)={\displaystyle \frac{\mathrm{\Gamma }}{2\pi }}{\displaystyle \frac{1}{(mm_0)^2+\frac{\mathrm{\Gamma }^2}{4}}}|\omega (m)|^2`$ (32) leads to the well known exponential decay law which follows from the survival amplitude $`A_{BW}(t)={\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑me^{imt}\rho _{WB}(m)=e^{im_0t}e^{\frac{1}{2}\mathrm{\Gamma }|t|}.`$ (33) (Note that the existence of the ground state induces non-exponential corrections to the decay law and to the survival amplitude (33) — see ). It is therefore reasonable to assume a suitable form for $`\omega _{S,\beta }`$ and $`\omega _{L,\beta }`$. More specifically, we write $$\omega _{S,\beta }(m)=\sqrt{\frac{\mathrm{\Gamma }_S}{2\pi }}\frac{A_{S,\beta }(K_S\beta )}{mm_S+i\frac{\mathrm{\Gamma }_S}{2}},\omega _{L,\beta }(m)=\sqrt{\frac{\mathrm{\Gamma }_L}{2\pi }}\frac{A_{L,\beta }(K_L\beta )}{mm_L+i\frac{\mathrm{\Gamma }_L}{2}}$$ (34) where $`A_{S,\beta }`$ and $`A_{L,\beta }`$ are decay (transition) amplitudes, end thus $$\rho _{x,\beta }(m)=\frac{\mathrm{\Gamma }_x}{2\pi }\frac{(A_{x,\beta }(K_x\beta ))^2}{(mm_x)^2+\frac{(\mathrm{\Gamma }_x)^2}{4}},$$ (35) where $`x=L,S`$. In the one-pole approximation (34) $`A_{K^0K^0}(t)`$ can be conveniently written as $`A_{K^0K^0}(t)`$ $`=`$ $`A_{\overline{K}^0\overline{K}^0}(t)=`$ (36) $`=`$ $`{\displaystyle \frac{1}{2\pi }}\{e^{im_St}({\displaystyle _0^{\frac{m_S}{\gamma _S}}}dy{\displaystyle \frac{e^{i\gamma _Sty}}{y^2+1}}+{\displaystyle _0^{\mathrm{}}}dy{\displaystyle \frac{e^{i\gamma _Sty}}{y^2+1}})+`$ $`+e^{im_Lt}({\displaystyle _0^{\frac{m_L}{\gamma _L}}}dy{\displaystyle \frac{e^{i\gamma _Lty}}{y^2+1}}+{\displaystyle _0^{\mathrm{}}}dy{\displaystyle \frac{e^{i\gamma _Lty}}{y^2+1}})\}.`$ Collecting only exponential terms in (36) one obtains an expression analogous to the WW approximation $$A_{K^0K^0}(t)=A_{\overline{K}^0\overline{K}^0}(t)=\frac{1}{2}\left(e^{im_St}e^{\gamma _St}+e^{im_Lt}e^{\gamma _Lt}\right)+N_{K^0K^0}(t).$$ (37) Here $`N_{K^0K^0}(t)`$ denotes all non-oscillatory terms present in the integral (36). ## 3 Diagonal matrix elements of the effective <br>Hamiltonian Using the decomposition of type (37) and the one-pole ansatz (34), we find the difference (25), which is now formulated for the $`K^0\overline{K}^0`$ system. Here it has the following form: $`h_{11}(t)h_{22}(t)={\displaystyle \frac{X(t)}{Y(t)}},`$ (38) where $`X(t)=i\left({\displaystyle \frac{A_{\overline{K}^0K^0}(t)}{t}}A_{K^0\overline{K}^0}(t){\displaystyle \frac{A_{K^0\overline{K}^0}(t)}{t}}A_{\overline{K}^0K^0}(t)\right)`$ (39) and $`Y(t)=A_{K^0K^0}(t)A_{\overline{K}^0\overline{K}^0}(t)A_{K^0\overline{K}^0}(t)A_{\overline{K}^0K^0}(t).`$ (40) Using the above mentioned spectral formulae in the one - pole approximation (34) we get $`A_{K^0\overline{K}^0}(t)`$ and $`A_{\overline{K}^0K^0}(t)`$ $`A_{K^0\overline{K}^0}(t)={\displaystyle \frac{1+\pi }{8\pi p^{}q}}\{e^{im_St}e^{\gamma _St}[1+`$ $`+{\displaystyle \frac{\sqrt{\gamma _S\gamma _L}}{\gamma _S}}(2i\gamma _SC_I+D_I^{}F_I^{})]+`$ $`+e^{im_Lt}e^{\gamma _Lt}[1+`$ $`+{\displaystyle \frac{\sqrt{\gamma _S\gamma _L}}{\gamma _L}}(2i\gamma _LC_ID_I^{}+F_I^{})]\}+`$ $`+N_{K^0\overline{K}^0}(t)`$ (41) and $`A_{\overline{K}^0K^0}(t)={\displaystyle \frac{1+\pi }{8\pi pq^{}}}\{e^{im_St}e^{\gamma _St}[1+`$ $`+{\displaystyle \frac{\sqrt{\gamma _S\gamma _L}}{\gamma _S}}(2i\gamma _SC_ID_I^{}+F_I^{})]+`$ $`+e^{im_Lt}e^{\gamma _Lt}[1+`$ $`+{\displaystyle \frac{\sqrt{\gamma _S\gamma _L}}{\gamma _L}}(2i\gamma _LC_I+D_I^{}F_I^{})]\}+`$ $`+N_{\overline{K}^0K^0}(t),`$ (42) where $`N_{K^0\overline{K}^0}(t)`$, $`N_{\overline{K}^0K^0}(t)`$ denotes all non-oscillatory terms and $`C_I,D_I^{},F_I^{}`$ are defined in . Using the expression for the derivative of $`E_i`$ we can find the derivatives which will be necessary for the following calculations $`\frac{A_{K^0\overline{K}^0}(t)}{t}`$ and $`\frac{A_{\overline{K}^0K^0}(t)}{t}:`$ $`{\displaystyle \frac{A_{K^0\overline{K}^0}(t)}{t}}`$ $`=`$ $`{\displaystyle \frac{1+\pi }{8\pi p^{}q}}\{e^{im_St}e^{\gamma _St}[im_S\gamma _S+`$ (43) $`+\sqrt{\gamma _S\gamma _L}(2i\gamma _SC_ID_I^{}+F_I^{})]+`$ $`+e^{im_Lt}e^{\gamma _Lt}[im_L\gamma _L+`$ $`+\sqrt{\gamma _S\gamma _L}(2i\gamma _LC_I+D_I^{}F_I^{})]\}+`$ $`+\mathrm{\Delta }N_{K^0\overline{K}^0}(t)`$ and $`{\displaystyle \frac{A_{\overline{K}^0K^0}(t)}{t}}`$ $`=`$ $`{\displaystyle \frac{1+\pi }{8\pi pq^{}}}\{e^{im_St}e^{\gamma _St}[im_S\gamma _S+`$ (44) $`+\sqrt{\gamma _S\gamma _L}(2i\gamma _SC_I+D_I^{}F_I^{})]+`$ $`+e^{im_Lt}e^{\gamma _Lt}[im_L\gamma _L+`$ $`+\sqrt{\gamma _S\gamma _L}(2i\gamma _LC_ID_I^{}+F_I^{})]\}+`$ $`+\mathrm{\Delta }N_{\overline{K}^0K^0}(t),`$ where $`\mathrm{\Delta }N_{K^0\overline{K}^0}(t)`$, $`\mathrm{\Delta }N_{\overline{K}^0K^0}(t)`$ denotes all non-oscillatory terms. The states $`|K_L`$ and $`|K_S`$ are superpositions of $`|K^0`$ and $`|\overline{K}^0`$. The lifetimes of particles $`|K_L`$ and $`|K_S`$ may be denoted by $`\tau _L`$ and $`\tau _S`$, respectively, $`\tau _L=\frac{1}{\gamma _L}=5,18310^8s`$ being much longer than $`\tau _S=\frac{1}{\gamma _S}=0,892310^{10}s.`$ Below we calculate the difference (38) for $`t\tau _L`$ $`h_{11}(t\tau _L)h_{22}(t\tau _L)={\displaystyle \frac{X(t\tau _L)}{Y(t\tau _L)}}.`$ (45) If we only consider the long living states $`|K_L`$ we may drop all the terms containing $`e^{\gamma _St}|_{t\tau _L}`$ as they are negligible in comparison with elements involving the factor $`e^{\gamma _Lt}|_{t\tau _L}.`$ We also drop all the non-oscillatory terms $`N_{K^0K^0}(t),`$ $`N_{\overline{K}^0K^0}(t)`$, $`N_{K^0\overline{K}^0}(t)`$ present in $`A_{K^0K^0}(t)`$, $`A_{\overline{K}^0K^0}(t)`$ and $`A_{K^0\overline{K}^0}(t),`$ that is in integrals (36), (41) and (42), because they are extremally small in the region of time $`t\tau _L`$ . Similarly, because of the properties of the exponential integral function $`E_i,`$ we can drop terms like $`\mathrm{\Delta }N_{\overline{K}^0K^0}`$ and $`\mathrm{\Delta }N_{K^0\overline{K}^0}`$ present in $`\frac{A_{\overline{K}^0K^0}}{t}`$ (43) and $`\frac{A_{K^0\overline{K}^0}}{t}`$ (44).This conclusion follows from the asymptotic properties of the exponential integral function $`E_i`$ and the fact that $`\mathrm{\Delta }N_{\overline{K}^0K^0},`$ $`\mathrm{\Delta }N_{K^0\overline{K}^0}`$ only contain expressions proportional to $`E_i`$. We may now calculate the products $`A_{K^0K^0}(t)A_{\overline{K}^0\overline{K}^0}(t),`$ $`A_{K^0\overline{K}^0}(t)A_{\overline{K}^0K^0}(t),`$ $`\frac{A_{\overline{K}^0K^0}}{t}(t)A_{K^0\overline{K}^0}(t),`$ $`\frac{A_{K^0\overline{K}^0}}{t}(t)A_{\overline{K}^0K^0}(t)`$, which, after using the above mentioned properties of $`N_{K^0K^0}(t)`$, $`\mathrm{\Delta }N_{K^0K^0}(t)`$ and performing some algebraic transformations, leads to the following form of the difference (45): $`h_{11}(t\tau _L)h_{22}(t\tau _L))=({\displaystyle \frac{2\pi ^2\sqrt{\gamma _S\gamma _L}}{\pi ^2+2\pi +1}}){\displaystyle \frac{Z}{W}}0,`$ (46) where $`Z=`$ $`4|p|^2|q|^2{\displaystyle \frac{\pi ^2+2\pi +1}{4\pi ^2}}[1+`$ (47) $`+\gamma _S(4\gamma _LC_I^2+{\displaystyle \frac{1}{\gamma _L}}(D_I^{}_{}{}^{}2F_I^{}_{}{}^{}2+4D_I^{^{}}F_I^{^{}})+`$ $`+4iC_I(D_I^{^{}}F_I^{^{}}))]0`$ $`W=`$ $`2(C_Im_L+D_I^{^{}}F_I^{^{}})+`$ (48) $`+i[4C_I\gamma _L+{\displaystyle \frac{m_L}{\gamma _L}}(D_I^{^{}}+F_I^{^{}})]0.`$ ## 4 Final remarks * Our results presented in the present poster have shown that in a CPT invariant and CP noninvariant system in the case of the exactly solvable one-pole model, the diagonal matrix elements do not have to be equal. In the general case the diagonal elements depend on time and their difference, for example at $`t\tau _L`$, is different from zero. Z and W in (46) are different from zero, so the difference $`(h_{11}(t)h_{22}(t))|_{t\tau _L}0.`$ From this observation a conclusion of major importance can be drawn, namely that the measurement of the mass difference $`(m_{K^0}m_{\overline{K}^0})`$ should not be used while designing CPT invariance tests. This runs counter to the general conclusions following from the Lee, Oehme and Yang theory. * A detailed analysis of $`h_{jk}(t)`$, $`(j,k=1,2)`$ shows that the non-oscillatory elements $`N_{\alpha ,\beta }(t),\mathrm{\Delta }N_{\alpha ,\beta }(t)`$ (where $`\alpha ,\beta =K^0,\overline{K}^0`$) is the source of the non-zero difference $`(h_{11}(t)h_{22}(t))`$ in the model considered. It is not difficult to verify that dropping all the terms of $`N_{\alpha ,\beta }(t),\mathrm{\Delta }N_{\alpha ,\beta }(t)`$ type in the formula for $`(h_{11}(t)h_{22}(t))`$ gives $`(h_{11}^{osc}(t)h_{22}^{osc}(t))=0`$, where $`h_{jj}^{osc}(t)`$, $`(j=1,2)`$, stands for $`h_{jj}(t)`$ without the non-oscillatory terms. * The result ($`h_{11}(t)h_{22}(t))0`$ seems to be very important as it has been obtained within the exactly solvable one-pole model based on the Breit-Wigner ansatz, i.e. the same model as used by Lee, Oehme and Yang. ## Acknowledgements The author wishes to thank Professor Krzysztof Urbanowski for many helpful discussions.
warning/0506/quant-ph0506045.html
ar5iv
text
# Generalized quantum measurements. Part I: Information properties of soft quantum measurements ## I Introduction One of the fundamental transformations in quantum physics is the projective measurement transformation, which sets a correspondence between quantum theory and real physical experiment landauqm ; mensky . Despite the fact that physical realization of this type of quantum transformations, which preserve the states of the measurable quantum variables, i.e., *nondemolition measurements*, causes experimental difficulties, this field of experimental quantum physics is still of prime importance Grangier1998 ; Pryde2004 . From theoretical point of view, the projective measurement in its simplest variant, i.e., with no addressing to its generalized variant of indirect measurement, sets one-to-one correspondence between the orthogonal, i.e., entirely distinguishable, set of states $`|k`$ of the measurable system (we will call it in the following simply the *object*) and the set of orthogonal states of the measuring device (we will call it in the following simply the *meter*) accompanied with a complete loss of the phase relationships between them. In this case, the principal point is that one can use for the measurement a *classical* meter neumann , which corresponds to the case, when in the final state of the object–meter system after the measuring process take part only those meter states for which their quantum or microscopical nature is not essential. Nowadays, however, progress in experimental quantum physics allows powerful tools for preparation and manipulation with the quantum states QC , so that the standard concept of the quantum measurement as the projective measurement transformation can be and has to be revised towards releasing the limitation by the quasiclassical meter only and generalizing the concept of the quantum measurement on the case when the meter is an essentially quantum device. Such generalization of the quantum measurement concept is, obviously, necessary for the adequate discussion of modern experiments in the field of engineering of quantum information using such objects as atoms and ions in various traps and photons of electromagnetic field. At the same time, the most useful is not the most possible generalized concept of the quantum measurement, as of an arbitrary quantum transformation, which contains information about the measurable variables (see, for instance, vedral02 ; ozawa01 ; barnum02 ), but selection of those transformations, which, similar to projective measurement, are based on one-to-one correspondence between the initial states of the object and final states of the meter. The mechanisms of mapping the quantum information via setting one-to-one correspondence between quantum events, i.e., with the help of the corresponding classical information index $`k`$, is undoubtedly the most important representation of quantum information relations and has fundamental value for understanding of the basic grounds of quantum mechanics. In this work, we consider the so called soft quantum measurement, which is the simplest model for a generalized quantum measurement that introduces in a concentrated form all essential physical mechanisms responsible for inevitable disturbance of the initial quantum information and the competitive character of the process of its attainment, which results, as it is widely known, in the possibility of reliable detection of the fact of its unauthorized usage. The paper is organized as follows. In Sec. II, the general class of the information-preserving measurements is defined. In Sec. III the definition and general discussion of the soft measurement is presented. In Sec. IV the general properties of the repeated soft measurements are discussed. Sec. V presents the quantitative analysis of the information properties of the soft measurements. Finally, the key results of the paper are summarized in conclusions. ## II Quantum measurements preserving information By definition, the mapping $`\psi \phi `$ of the input states $`\psi `$ into the output states $`\phi `$ is represented by the projection operator $`|\phi \psi |`$ or a *pointer*. For the measurement, we do understand under $`\psi `$ only quantum states of the object, whereas $`\phi `$ at the minimal complete description must represent all essential under the measurement the object–meter quantum states in the respected Hilbert space $`H_AH_B`$, which, in the general case, must be supplemented with the reservoir $`H_D`$. Obviously, the most direct realization of the classical measurement concept is the transformation that is described with the set of projectors $`|\phi \psi |=||kk|`$ with the classical index $`k`$ that numbers the input quantum information. The content of the input information allows establishment of such correspondence with the classical parameter due to the assumed orthogonality of the input states $`|k`$ of the object. The latter form the orthogonal basis of the space $`H_A`$ and are the eigen-states of a measurable physical variable $`\widehat{A}=\lambda _k|kk|`$. For the measurement superoperator representation in the form of the Kraus expansion $`_k\widehat{F}_k\widehat{F}_k^+`$, where $``$ is the substitution symbol for the transformed density matrix, it corresponds to the operators $`\widehat{F}_k=||kk|`$. The nondemolition projective measurement corresponds to the choice $`||k=|k|k`$, which realizes one-to-one correspondence between the initial and the resulting bases of the measurable object, which coincide, and those of the meter at the complete loss of the phase relationships between their elements. A simplest generalization of the projective measurement, which takes into account essentially quantum character of the meter in the frame of the concept of the “ideal”, i.e., nondemolition and absolutely precise, measurement is given with the concept of the so called *entangling measurement* PRA2003 ; OptSpectr2004 . Following this concept, the entangling measurement presents the measured information in the same form of 100%-correlations, i.e., as one-to-one correspondence between the states of the object and the meter but, in general case, does not destroy the coherency of the measurable states $`|k`$ completely. Instead, this measurement represents the resulted information in the form of quantum entanglement in the bipartite object–meter system leaving the measurable states $`|k`$ unperturbed harocheRef . The entangling measurement maps the initial states $`\psi =c_k|k`$ of the object onto the resulted states $`\psi _{ABD}=c_k|k|k\left|k\right)`$ of the tripartite system $`A+B+D`$ object–meter–reservoir, where $`|k`$, $`|k`$ are the orthogonal bases in $`H_A`$ and $`H_B`$, respectively; $`\left|k\right)`$ is the set of (not necessarily orthogonal) states of the reservoir with the scalar product $`R_{kl}=(k|l)`$, which is the so called *entanglement matrix* that represents the respective dephasing effects in the object–meter system SPIE2004i ; LLasPhys . The corresponding mapping of the initial state of the object onto the quantum states of the bipartite system object–meter are given by the superoperator $$_e=\underset{kl}{}R_{kl}|k|kl|l|k||l.$$ (1) Here the transformed density matrix $`\widehat{\rho }_A`$ is determined in the Hilbert space of the object $`H_A`$, whereas the transformation result $`\widehat{\rho }_{AB}`$ is determined in the bipartite space $`H_AH_B`$. Eq. (1) does not include the initial state of the object because we assume it either specified *a priori* or an arbitrary one, but forgettable at the process of measurement. This initial state does not affect the transfer of the measurable information. At $`R_{kl}=\delta _{kl}`$, the considered superoperator describes the standard projective measurement. At $`R_{kl}1`$, we have the so called *pre-measurement* schlosshauer , which corresponds to the total preservation of the initial coherency and transfer of all essentially quantum (coherent) information, initially stored in the object, onto the set of duplicated states $`|k|k`$ PRA00 . In the latter case, we have exact cloning of the orthogonal set of the measurable states and the superoperator (1) has the form $`_e=VV^+`$, where $`V=|k|kk|`$. In terms of wave functions, it corresponds to the isometric transformation $`V`$ from $`H_A`$ onto $`H_AH_B`$. By contrast with the standard classical measurement, in the entangling measurement the meter contains the exact value of the classical variable $`k`$, but due to the entanglement in the object-meter system this information cannot be red out and, specifically, be copied with the help of other physical systems without respective losses of the coherent information, which is created during the measurement. Without any loss, it is capable only in the case of the standard projective measurement corresponding to a purely incoherent set of the duplicated states $`|k|k`$. Preserving the measured basis states in their initial form, which is the essence of the nondemolition measurements, makes the entangling measurement a very specific transformation realization of which (as well as the nondemolition projective measurement) requires special efforts. From the information content of the resulting state of the object–meter system, the nondemolition character of the measurement gives no additional advantages, but defines how the coherent quantum information in the output is linked with the initial set of the object states $`|k`$. One can generalize the entangling measurements in a natural way by giving up the demand of nondestructiveness of the measurable states and, additionally, the demand of one-to-one correspondence of the states of the system and the meter. For this, in Eq. (1) the exactly cloned orthogonal states $`|k`$ of the object and the meter can be replaced with the nonorthogonal states $`\left|k\right)`$, which contain the internal indeterminacy and cannot be cloned, in principle. After such a replacement the superoperator (1) remains positively defined, but, as one can easily see, in order to preserve its normalization it is necessary and sufficient to fulfill the condition $`R_{kl}Q_{kl}^AQ_{kl}^B=0`$ for all $`k,l`$, where $`Q_{kl}^{A,B}=(k|l)^{A,B}`$ are the respective Gram matrix for the set of object and meter states. This means that the orthogonality for the given $`k,l`$ must be hold true at least in one of the subsystems of the object–meter–reservoir system, which is due to the unitarity of the mapping, considered in the terms of the complete system evolution. Respectively, for the completely coherent measurement, $`R_{kl}1`$, the possibility of using nonorthogonal resulting states of the object and the meter has an alternative character, i.e., for the output object-system states we have either $$(I)||k=|k\left|k\right)\text{or}(II)||k=\left|k\right)|k.$$ (2) The first considered possibility corresponds to the case of a nondemolition “fuzzy” measurement, which confronts the distinct measurable states with not entirely distinguishable states of the meter. The second possibility corresponds to a particular case of an exact but “destructive” measurement, which changes the initial basis states of the object. Both these classes of measurements are the measurements that completely *preserve information* in the joint states of the object–meter system about the measurable states of the object regardless to the presence or absence of the external dephasing (reservoir). All the measurements considered in this paper, including those defined with the superoperator (1), belong to the general class of *nondemolition measurements* that preserve the complete set of the classically compatible object states $`|k`$, $`_k|k=H_A`$. The most general transformation for this class of measurements is described by the superoperator of the form $$_{\mathrm{nd}}=\underset{kl}{}\left(|kl|\widehat{\rho }_{kl}^M\right)k||l,$$ (3) where the set of operators $`\widehat{\rho }_{kl}^M`$ in Hilbert space $`H_B`$, which describes essential for the measurement variables of the meter, defines the positive block-type operator with the normalized diagonal terms, $`\mathrm{Tr}\widehat{\rho }_{kk}^M=1`$. This superoperator associates the object projectors $`|kl|`$ with the $`kl`$-elements of the block-type operator $`\left(\widehat{\rho }_{kl}^M\right)=_{kl}|kl|\widehat{\rho }_{kl}^M`$ in the object–meter system, which, in general case, describes the states of the meter entangled with the measurable states of the object. Here, normalization of the diagonal terms ensures preserving of the probability for the set of compatible measurable states of the object $`|kk|`$, which are not perturbed during the measurement. For the trivial case of $`\widehat{\rho }_{kl}^M=\widehat{\rho }_0^B`$, $`_{\mathrm{nd}}\widehat{\rho }^A=\widehat{\rho }^A\widehat{\rho }_0^B`$, all states of the meter are associated with a single density matrix of the meter states, i.e., performing no measurement. In case of the entangling measurement (1), we have a set $`\widehat{\rho }_{kl}^M=R_{kl}|kl|`$, which establishes correspondence one-to-one between measurable states of the object with the similar, orthogonal and, respectively, completely distinguishable states of the meter. Such a measurement is the distinct one in the sense that for the compatible states of the object and meter $`\widehat{\rho }^{AB}=_{\mathrm{nd}}\widehat{\rho }^A`$ the joint probability distribution for the respective events $`\widehat{P}_k^A=|k|k\widehat{I}_B`$, $`\widehat{P}_l^B=\widehat{I}_A|l|l`$ is singular: $`P(k,l)=\mathrm{Tr}\widehat{P}_k^A\widehat{P}_l^B\widehat{\rho }^{AB}=\delta _{kl}\rho _{kk}^A`$, i.e. in the supporting subspace of the density matrix $`\widehat{\rho }^{AB}`$ we have $`\widehat{P}_k^A=\widehat{P}_k^B`$ for all $`k`$. ## III Definition and physical essence of the soft measurement We consider here the most fundamental class of quantum nondemolition measurements, which is characterized by using an *“unclear”* set of nonorthogonal states $`\left|k\right)`$ that contain the internal quantum uncertainty for indication of the measurements results. This kind of measurement in the limiting cases of the orthogonal or the trivial (consisted of the only state $`\left|k\right)|0`$) sets is reduced to the described above entangling (specifically, to projective) measurement and to the no measurement, respectively. In the latter case, the only reason for the changes in the object–meter system is the interaction of the object with the dephasing subsystem, which takes place at $`R_{kl}1`$. In a general case, the measurements of this type, which just slightly change the initial state of the object, are usually called the *fuzzy* measurements mensky ; schlosshauer . For this generalized measurement, the meter does not contain any specified physical variable, which can store exact information about the number $`k`$ of the measurable object states $`|k`$, and the attained information is connected with the entire physical structure of the meter and is represented in essentially quantum form. Note that the most comprehensive description of the measurement is related to the consideration of the complete object–meter system, whereas the quantum analysis of the measurement as the transformation only in the space of the object states, which is performed in many papers, does not reflect all essentially quantum information in the object–system system. We will call the measurement (3) the soft measurement, when the resulting information of the meter is reflected in the matrix elements $`\widehat{\rho }_{kl}^M=R_{kl}\left|k\right)\left(l\right|`$ by pure states $`\left|k\right)H_B`$, uncertainty of which has purely quantum nature and connected with their nonorthogonality, which leads to the impossibility of setting one-to-one correspondence between the measurable orthogonal states similar to the classically distinguishable states of the meter. Such a measurement, described with the superoperator $$=\underset{kl}{}R_{kl}|k\left|k\right)k||l\left(l\right|l|,(k|l)=Q_{kl},$$ (4) sets one-to-one correspondence between the object states $`|k`$ and those of the meter, $`\left|k\right)`$. Its physical realizability is ensured by the complete positivity kraus of the transformation (4) for $`R0`$. Physical essence of this transformation reduces to the independent from the initial meter state transformation of the initial orthogonal basis states of the object $`|k`$ into also orthogonal, i.e., completely distinguishable, states $`|k\left|k\right)`$ of the bipartite object–meter system. At the same time, the phase relationships between the initial states are generally perturbed and their joint correlations are described with the matrix elements $`R_{kl}`$, whereas the Gram matrix $`Q`$ describes the degree of quantum distinguishability of the measuring states $`\left|k\right)`$ of the meter. The soft character of the measurement is reflected by the difference of matrix $`Q`$ from the identity matrix $`Q=I`$, which corresponds to the conventional (distinct) entangling measurement. For the transformation (4), as well as in the case of distinct measurement, the classical content of the measurable object states $`|k`$ is not perturbed, whereas the quantum information initially stored in the initial object state $`\widehat{\rho }^A`$ is redistributed between two subsystems and perturbed due to dephasing. In the limiting case of the trivial set $`\left|k\right)|0`$ the superoperator (4) corresponds to the independent from the meter transformation of the object with partial loss of the joint coherency of the measurable states $`|k`$. With this, the coherency of the object is preserved in case of $`R_{kl}=e^{i(\phi _k\phi _l)}`$, i.e., with the determinate phase transformation. <sup>1</sup><sup>1</sup>1The given density matrix is positively determined, which is necessary and sufficient for the complete positiveness of the superoperator. Note that deviation from the phases difference $`\phi _k`$ in the exponent violates the positiveness. When the coherency is preserved, the soft measurement, considered as the transformation in the bipartite object–meter system with the initially “prepared” pure state $`|0`$ of the meter, is equivalent to the unitary transformation. It maps the set of initial orthogonal states of the form $`|k|0`$ onto the orthogonal states $`\left|k\right)\left|k\right)`$ and, obviously, can be redefined up to the unitary operator $`U_{AB}`$ in the total space $`H_AH_B`$. The respective redefinition of the superoperator (4) then can be represented with the superposition $`𝒰𝒮`$ of the superoperator $`𝒮=|00|\mathrm{Tr}_B`$ of the resetting the meter into the initial state $`|0`$ and of the unitary superoperator transformation $`𝒰=U_{AB}U_{AB}^1`$. One can also easily see that the entropy of the initial object state is entirely transferred into the entropy of the bipartite object–meter system, $`S[\widehat{\rho }^A]=S[\widehat{\rho }^{AB}]`$. Respectively, the coherent information schumacher , defined with respect to its transformation from $`H_A`$ into $`H_AH_B`$ PRA00 , is equal to its initial value $`S[\widehat{\rho }^A]`$. All the losses are due to only the dephasing and the respective violation of the isometricity of the transformation at $`R_{kl}1`$. It is not difficult to calculate now the Hamiltonian of the transformation of the infinitesimal fuzzy measurement in the object–meter system with the fixed initial state $`|0`$ of the meter, which can be chosen as one of the resulting states of the meter, i.e., $`|0=\left|0\right)`$. Calculating the infinitesimal addition for a short time in the state of the object–meter system as the result of the corresponding unitary transformation with the “unperturbed” or “perturbation-free” Hamiltonian $`\widehat{\epsilon }=_k|kk|\widehat{\epsilon }_B(k)`$ and equating its result to the change, which is caused by the transformation (4), we do have $`i\frac{\mathrm{\Delta }t}{\mathrm{}}\widehat{\epsilon }|k|0=|k\left|\delta k\right),`$ where $`\left|\delta k\right)=\left|k\right)|0`$. From here, for the $`k`$-dependent Hamiltonian of the meter we receive $`i\frac{\mathrm{\Delta }t}{\mathrm{}}l|\widehat{\epsilon }_B(k)|0=l|\delta k)`$, from which, due to the hermicity, follows the equation for the uniquely determined matrix elements: $$\widehat{\epsilon }_B(k)=\underset{l}{}\underset{t0}{lim}i\frac{\mathrm{}}{\mathrm{\Delta }t}(l|\delta k)|l0|(\delta k|l|0l|),$$ (5) whereas other elements can be defined arbitrary way or, for example, be set to zero. In the latter case, Eq. (5) has the structure $`i(\widehat{a}\widehat{a}^+)`$. Generalization to the case when external dephasing is present can be described with the unitary transformation in the system, which contains an additional dephasing degree of freedom $`H_D`$, that, by analogy with the transformation introduced in Ref. LLasPhys, for the case of the entangling measurement, includes also the infinitesimal unitary transformation created by the Hamiltonian (5). ## IV Repeated measurements ### IV.1 Measurements at the output of the meter The repeated application of the fuzzy measurement to the *result* of the initial measurement does not increase the attained information because the resulted information contains an additional indeterminacy in comparison with a single measurement, which does not vanish or decreases at the second (or repeated) interaction of the meter with the object. ### IV.2 Repeated measurements of the object with the accumulation of the information For the repeated measurement of the *object* with preservation of the measurement results in independent degrees of freedom of the multicomponent meter due to the $`n`$-fold application of the measurement transformation we receive the following resulting transformation: $$^{(n)}=\underset{kl}{}R_{kl}^n\left|k\right)\mathrm{}\left|k\right)|kk||ll|\left(l\right|\mathrm{}\left(l\right|.$$ (6) This transformation results in increasing incoherency and yields the multiply duplicated unclear information about the value $`k`$. At the same time, the quantum character of the measurable information is maximally preserved only in the entire system (that includes all the meter’s subsystems). After the averaging over the $`m<n`$ output subsystems of the meter the remaining quantum information is dephased and is characterized with the entanglement matrix $`R_{kl}^{(nm)}=Q_{kl}^mR_{kl}^n`$. This matrix defines the incoherency of the measurement even without any dephasing during creation of the entanglement in separate measurements, i.e., for $`R_{kl}1`$. In active subspace $`H_D=\mathrm{sp}\{||k,k=1,\mathrm{},D\}H_B^n`$ of the *collective* states $`||k=\left|k\right)\mathrm{}\left|k\right)`$ the measurement transformation (6) has the form of a single measurement, but with entanglement matrices and scalar products corresponding to $`n`$ measurements. Such measurement can be illustrated with many physical realizations. For instance, if we select an atom from an atomic gas (ensemble) as a quantum object, we can consider, in general, all the surrounding atoms as the multipartite meter system. Then, separate collisions can be considered as the separate measurement acts, which augments the measurable information in the multipartite system. This situation, surely, exceeds the bounds of the standard quantum measurement, which assures that all the measurable information is accessible and can be used for any purpose. Another physical situation with the atoms trapped in an optical dipole trap granjier ; meschede fits our model of the repeated measurements more precisely. In an optical dipole trap, an atom moving along the trapped in the micropotential holes atoms performs repeated measurements (Fig. 1). Successful experimental realizations of the nondemolition projective measurements with a single photon haroche ; Pryde2004 give us a hope that the repeated measurements considered here would be realized experimentally in the nearest future not only with atoms, but also with photons. The dimension of the Hilbert space of the active states of the meter does not exceed the dimension of the space with the entire set of states of the meter, despite the fact that the space of the meter’s states unrestrictedly expands. For appearance of new active states different from the set $`|k\left|k\right)\mathrm{}\left|k\right)`$ (or to the unitary equivalent to it) it is necessary that the dynamics along the different degrees of freedom be independent and random. However, deviation of the bases used in different measurements leads to the resulting states of the meter from the indicated active space and if we have no *a priori* information about these deviations will result in losses of information about the object. Therefore, coding of information in the series of repeated fuzzy measurements provides a resource for the latent storing of information with the quantum key, which is unique. When all the measurement results are preserved, the joint density matrix of the object–meter system can be written as $$\widehat{\rho }^{B^nA}=\underset{kl}{}R_{kl}^n\rho _{kl}^A\left|k\right)\mathrm{}\left|k\right)|kl|\left(l\right|\mathrm{}\left(l\right|.$$ (7) In its turn, averaged over the states of the object density matrix of the meter is $$\widehat{\rho }^{B^n}=\rho _{kk}^A\left|k\right)\mathrm{}\left|k\right)\left(k\right|\mathrm{}\left(k\right|,$$ (8) i.e., represented with the weighted sum of $`D`$ non-commuting projectors. In orthogonal basis $`||e_k`$ of the active subspace of the collective states $`H_D`$, they can be rewritten in the form $`\widehat{\rho }^{B^nA}`$ $`=`$ $`{\displaystyle \rho _{ki,lj}^{B^nA}||e_k|ij|e_l||},`$ (9) $`\widehat{\rho }^{B^n}`$ $`=`$ $`{\displaystyle \rho _{kl}^{B^n}||e_ke_l||}`$ (10) with matrix elements corresponding to the respective Eqs. (7) and (8) and the choice of the basis $`||e_k`$. The matrix of scalar products for the vectors set $`\left|k\right)\mathrm{}\left|k\right)`$ has the form $`Q^{(n)}=\left(Q_{kl}^n\right)`$. In the case of linearly independent set $`\left|k\right)`$ for $`n\mathrm{}`$ this matrix has the form of identity matrix $`Q^{(\mathrm{})}=I`$. The orthonormalized basis in $`H_D`$ can be expressed via the duplicated states of the meter with the help of the following relationship: $$||e_k=\underset{l}{}\left(Q_{}^{(n)}{}_{}{}^{1/2}\right)_{kl}^{}\left|l\right)\mathrm{}\left|l\right).$$ (11) Here the corresponding formulas $`\rho _{ki,lj}^{B^nA}`$ $`=`$ $`e_k||i|\widehat{\rho }^{B^nA}|j||e_l=R_{ij}^n\rho _{ij}^A\psi _i^{B^n}(k)\psi _j^{B^n}(l),`$ (12) $`\rho _{kl}^{B^n}`$ $`=`$ $`e_k||\widehat{\rho }^{B^n}||e_l={\displaystyle \underset{j}{}}\rho _{jj}^A\psi _j^{B^n}(k)\psi _j^{B^n}(l)`$ (13) represent the states of the meter with the matrix $$\psi _i^{B^n}(k)=\left(Q_{}^{(n)}{}_{}{}^{1/2}\right)_{ki}$$ (14) of fixed dimension $`D\times D`$, the latter (to avoid confusion of the “dimension” and “matrix”!) does not depend on the number of measurements $`n`$ and respective total dimension $`D^n`$ of the multipartite Hilbert space $`H_B^n`$ of the meter. The entanglement matrix $`R_{ij}`$ is essential only for constructing of the bipartite density matrix of the object–meter system and does not affect the partial density matrix of the meter because after the tracing over the object states their coherence is not important. A set of non-orthogonal, in general case, functions $`\psi _i^{B^n}(k)`$ makes sense of the ensemble of pure collective states of the meter, post-selected after $`n`$ measurements and corresponding to the $`i`$-th measured object states. All of them satisfy the normality condition $`(\psi _i^{B^n},\psi _i^{B^n})1`$, in which, due to definition (14) of collective states, the scalar products reduce to the normalized on unit the diagonal elements of the matrix $`Q^{(n)}`$. ### IV.3 Two-level system Let us choose the representation in which $`|0=(1,0)`$, $`|1=(0,1)`$, and the unclear set of measurable states has the form: $$\left|0\right)=(1,0),\left|1\right)=e^{i\chi }(\mathrm{cos}\frac{\vartheta }{2},e^{i\phi }\mathrm{sin}\frac{\vartheta }{2}).$$ (15) Then, matrices of the scalar products $$Q=\left(\begin{array}{cc}1& e^{i\chi }\mathrm{cos}\frac{\vartheta }{2}\\ e^{i\chi }\mathrm{cos}\frac{\vartheta }{2}& 1\end{array}\right);Q^{(n)}=\left(\begin{array}{cc}1& e^{in\chi }\left(\mathrm{cos}\frac{\vartheta }{2}\right)^n\\ e^{in\chi }\left(\mathrm{cos}\frac{\vartheta }{2}\right)^n& 1\end{array}\right),$$ depend only on the angle $`\vartheta `$ between the vectors $`\left|1\right)`$, $`\left|2\right)`$ and on its differential phase $`\chi `$. Applying relations (5) to the two-level case we have $$\widehat{\epsilon }_B(0)=0,\widehat{\epsilon }_B(1)=\underset{\mathrm{\Delta }t0}{lim}\frac{\mathrm{}}{\mathrm{\Delta }t}\left(\begin{array}{cc}2\mathrm{sin}\chi \mathrm{cos}\frac{\vartheta }{2}& ie^{i(\chi +\phi )}\mathrm{sin}\frac{\vartheta }{2}\\ ie^{i(\chi +\phi )}\mathrm{sin}\frac{\vartheta }{2}& 0\end{array}\right)=\mathrm{}\left(\begin{array}{cc}2\dot{\chi }& i\dot{\vartheta }\\ i\dot{\vartheta }& 0\end{array}\right),$$ (16) where $`\dot{\chi }`$, $`\dot{\vartheta }`$ describe the rates of changes of the respective angles in the process of the *only* measurement. We do not take into account here dependence on the second phase $`\phi `$, which describes the freedom in the choice of the common phase for the states $`\left|0\right)`$, $`\left|1\right)`$. For the matrix of wave functions of the measurable ensemble we have after $`n`$ measurements: $$Q_{}^{(n)}{}_{}{}^{1/2}=\frac{1}{2}\left(\begin{array}{cc}\sqrt{1\left(\mathrm{cos}\frac{\vartheta }{2}\right)^n}+\sqrt{1+\left(\mathrm{cos}\frac{\vartheta }{2}\right)^n}& e^{in\chi }\left[\sqrt{1+\left(\mathrm{cos}\frac{\vartheta }{2}\right)^n}\sqrt{1\left(\mathrm{cos}\frac{\vartheta }{2}\right)^n}\right]\\ e^{in\chi }\left[\sqrt{1+\left(\mathrm{cos}\frac{\vartheta }{2}\right)^n}\sqrt{1\left(\mathrm{cos}\frac{\vartheta }{2}\right)^n}\right]& \sqrt{1\left(\mathrm{cos}\frac{\vartheta }{2}\right)^n}+\sqrt{1+\left(\mathrm{cos}\frac{\vartheta }{2}\right)^n}\end{array}\right).$$ (17) Columns of this matrix play, in accordance with Eq. (14), role of the wave functions describing $`n`$-fold excitations of the meter in the minimal basis of the two-dimensional ($`D=2`$) space of the collective states. Let us consider a sequence of $`n`$ identical measurements, performed with the time period $`T`$ with small variation $`\vartheta `$ during the period time. We will also assume that the relationships for the parameters necessary for the asymptotically continuous changes of the result of the $`n`$-fold measurement as a function of the continuous time $`t`$ at $`nt`$. Respective continuous dynamics has a quantum diffusion character, which shows at short times, plus to the usual for the classical diffusion quadratic diffusional change of the state, a specific linear diffusional change in the non-diagonal matrix elements. Fluctuation character of the dynamics, which is typical for the description of the classical diffusion process with the use of stochastic equations, becomes apparent while considering the sequences of the respective classically compatible variables in the space $`H_B`$, which are in charge for the separate measurements in the whole measurement sequence. Their statistical description cannot be reduced to a reversible dynamics of the collective variables of the meter. Matrix (17) at $`\vartheta ^2=4\varkappa T0`$, $`n=t/T\mathrm{}`$, $`n\vartheta ^2=4\varkappa t=\mathrm{const}`$, and $`\chi =\dot{\chi }T`$ has the finite limit corresponding to the diffusion dynamics: $$Q^{1/2}(t)=\left(\begin{array}{cc}s_+(t)& e^{i\dot{\chi }t}s_{}(t)\\ e^{i\dot{\chi }t}s_{}(t)& s_+(t)\end{array}\right),$$ (18) where $`s_+(t)=\frac{1}{2}\left(\sqrt{1+e^{\varkappa t}}+\sqrt{1e^{\varkappa t}}\right)`$, $`s_{}(t)=\frac{1}{2}\left(\sqrt{1+e^{\varkappa t}}\sqrt{1e^{\varkappa t}}\right)`$. A similar limit has the entangling matrix $$R^{(n)}=(R_{ij}^n)=\left(\begin{array}{cc}1& r^n\\ r^n& 1\end{array}\right)\left(\begin{array}{cc}1& e^{\dot{r}t}\\ e^{\dot{r}^{}t}& 1\end{array}\right),\dot{r}=\underset{T0}{lim}\frac{1r}{T}.$$ This matrix describes the dequantization of the measurement result due to the external dephasing of the meter (reservoir). Simultaneously, at the long time, asymptotic diagonalization of the collective states of the meter occurs in accordance with the following asymptotic expression: $$Q^{1/2}(t)\stackrel{t\mathrm{}}{}\left(\begin{array}{cc}1\frac{1}{8}e^{2\varkappa t}& \frac{1}{2}e^{\varkappa t+i\dot{\chi }t}\\ \frac{1}{2}e^{\varkappa ti\dot{\chi }t}& 1\frac{1}{8}e^{2\varkappa t}\end{array}\right).$$ this matrix describes how the soft measurement transforms into the distinct completely coherent measurement (“pre-measurement”) with the orthogonal set of collective states of the meter $`||k|k`$. #### Partial density matrix of the meter From Eqs. (8), (18) we receive the partial density matrix of the meter in the process of continuous measurement in the form: $$\widehat{\rho }^{B^n}=\left(\begin{array}{cc}\frac{1}{2}+\frac{1}{2}\sqrt{1e^{\varkappa t}}\left(\rho _{11}^A\rho _{22}^A\right)& \frac{1}{2}e^{\varkappa t+i\dot{\chi }t}\\ \frac{1}{2}e^{\varkappa ti\dot{\chi }t}& \frac{1}{2}\frac{1}{2}\sqrt{1e^{\varkappa t}}\left(\rho _{11}^A\rho _{22}^A\right)\end{array}\right).$$ (19) At $`t=0`$, i.e., with no measurement, when for any initial states the resulted state has the only one and the same value, $`\left|2\right)=\left|1\right)=|0`$, the density matrix of the meter does not depend on $`\widehat{\rho }^A`$ and is equal to the projector $`|00|`$ onto $`|0`$, which in the given basis (11) for $`n=1`$ has the form $`|0=(1/\sqrt{2},1/\sqrt{2})`$. The choice of the basis states of the meter, which determines in the structure of the measurement superoperator (4) how the measured information is represented, is arbitrary. At $`t=\mathrm{}`$, i.e., in the limit of infinitive series of the limiting continuous soft measurements, the matrix (19) becomes a diagonal one with the matrix elements $`\rho _{11}^A`$, $`\rho _{22}^A`$, coinciding with the respective matrix elements of the measurable object. This coincidence holds not only for the limiting continuous measurements, but also for the series of measurements with the finite accuracy about the object states $`|k`$, which is determined with the tensor product of $`n`$ states $`\left|k\right)\mathrm{}\left|k\right)`$ in the form of multi-particle excitations of the $`n`$-fold copy of the measurable systems, the information in the limit $`n\mathrm{}`$ is retrievable. This leads to diagonalization of the marginal density matrix of the meter, which is constructed of the duplicated states of the object–meter system $`|k\left|k\right)\mathrm{}\left|k\right)`$, and to its coincidence with the density matrix of the object. The described exact coincidence of the representations $`\widehat{\rho }^{B^n}\widehat{\rho }^A`$ is due to the choice (11) of the basis linked to the object variable $`k`$. The entropy of the quantum state of the meter in the process of continuous measurement is changed from zero up to the entropy of the object. At the same time, the absence of the internal indeterminacy of the initial state of the meter is due to the absence of the information about the structure of the meter, which is reflected in its partial density matrix in the non-coherent, i.e., quasiclassical, form. #### Joint density matrix of the object–meter system The set of the orthogonal states of the object–meter system $`|k\left|k\right)\mathrm{}\left|k\right)`$ in the basis $`|l||e_m`$ in accordance with Eqs. (7), (12), and (18), where $`\psi ^{B^nA}`$ are supplemented with the basis object states, has the form: $$\psi ^{B^nA}=\{\begin{array}{c}s_+(t)||e_1|1+e^{i\dot{\chi }t}s_{}(t)||e_2|1,\\ e^{i\dot{\chi }t}s_{}(t)||e_1|2+s_+(t)||e_2|2.\end{array}$$ (20) Using this set and in accordance with Eq. (7), (12) the density matrix of the object–meter system can be written in the form of the $`4\times 4`$-matrix $$\widehat{\rho }^{B^nA}(t)=\left(\begin{array}{cc}\rho _{11}^A\left(\begin{array}{cc}s_+^2& s_+s_{}e^{i\dot{\chi }t}\\ s_+s_{}e^{i\dot{\chi }t}& s_{}^2\end{array}\right)& \rho _{12}^Ae^{\dot{r}t}\left(\begin{array}{cc}s_+s_{}e^{i\dot{\chi }t}& s_+^2\\ s_{}^2e^{2i\dot{\chi }t}& s_+s_{}e^{i\dot{\chi }t}\end{array}\right)\\ \rho _{21}^Ae^{\dot{r}^{}t}\left(\begin{array}{cc}s_+s_{}e^{i\dot{\chi }t}& s_{}^2e^{2i\dot{\chi }t}\\ s_+^2& s_+s_{}e^{i\dot{\chi }t}\end{array}\right)& \rho _{22}^A\left(\begin{array}{cc}s_{}^2& s_+s_{}e^{i\dot{\chi }t}\\ s_+s_{}e^{i\dot{\chi }t}& s_+^2\end{array}\right)\end{array}\right).$$ (21) ## V Information relationship in the processes of quantum measurements ### V.1 Coherent information in the object–meter and object–object channels Key features of the coherent information schumacher exchange at the entangling measurement (1) are described in detail in Ref. PRA2003, . The coherent information is purely quantum Barnum01 ; SPIE2001a ) and, therefore, it cannot be copied or duplicated. Thus, when the states $`|k`$ are duplicated, the coherent information is transferred onto the superposition of bipartite states $`|k|k`$ and lacking entirely in the channels “initial–resulting state of the object” and “initial state of the object–resulting state of the meter”. In accordance with its definition, the coherent information preserved in the channel, which realizes the superoperator transformation $`𝒩`$, can be written as $$I_c=S[𝒩\widehat{\rho }]S[\left(𝒩\right)\mathrm{\Psi }\mathrm{\Psi }^+],$$ (22) where the first term describes the entropy at the output of the channel and the second term—the so called exchange entropy that characterizes the entropy surge due to the irreversibility of the transformation; the pure state $`\mathrm{\Psi }`$ describes the so called “purified state” at the input of the channel as the state of the bipartite system input+reference, which describes the mixed input state as the result of its tracing over the auxiliary reference system, and $``$ is the identical transformation on the reference system state, which is not perturbed. Let us consider how the soft property of the measurement affects the transformation of the coherent information. In the two-time channel “object$``$object+meter”, the soft measurement does not affect the coherent information because of the orthogonality of the states $`|k\left|k\right)`$ of the system object–meter. All losses of the coherent information in this channel are due to the external dephasing only. By contrast, in the two-time channel “object–object” the influence of the soft measurement on the coherent information is a nontrivial one—in this channel the amount of preserved coherent information depends on both the external dephasing and the soft character of the measurement. The latter determines how information is distributed between the object and the meter. Substituting in Eq. (22) transformation $`𝒩=\mathrm{Tr}_B`$ and taking into account Eq. (4), we receive $`I_c=S\left[\left(R_{kl}Q_{kl}\rho _{kl}^A\right)\right]S\left[(R_{kl}Q_{kl}\mathrm{\Psi }_{ki}\mathrm{\Psi }_{lj}^{})\right]`$, where in brackets ($``$) are shown the matrix elements $`\stackrel{~}{\rho }_{kl}`$ and $`\stackrel{~}{\rho }_{ki,lj}`$ corresponding to the transformed density matrices of the object and object–reference system; $`_i\mathrm{\Psi }_{ki}\mathrm{\Psi }_{li}^{}=\rho _{kl}^A`$. For this channel, therefore, contributions from the external dephasing and distinctness of the measurement, which are presented with the respected matrices $`R_{kl}`$ and $`Q_{kl}`$, are totally equivalent. Simplifying the expression in the argument of the second term, one can rewrite the expression in the final form: $$I_c=S\left[\left(R_{kl}Q_{kl}\rho _{kl}^A\right)\right]S\left[\left(\sqrt{\rho _{kk}^A}R_{kl}Q_{kl}\sqrt{\rho _{ll}^A}\right)\right].$$ (23) In the absence of dephasing and at the maximal softness degree, i.e., $`|k|0`$, the second term in Eq. (23) vanishes, as far as its argument comes to a density matrix of a pure state, whereas the first term coincides with the entropy of the output, i.e., the coherent information is transmitted without disturbance to the object system only. In the opposite case, for the measurement with the complete distinguishability of the states of the meter or for their maximal external dephasing $`R_{kl}Q_{kl}=\delta _{kl}`$, both terms describe the entropy of the set of the measurable states of the object $`|k`$, which is determined with the maximum entropy probability distribution $`p_k=\rho _{kk}^A`$, and, respectively, the coherent information vanishes due to the complete dequantization of the input information. Calculation of the coherent information for a two-level system gives us the following expression: $$I_c=\frac{1}{2}[(1x_1)\mathrm{log}_2(1x_1)+(1+x_1)\mathrm{log}_2(1+x_1)(1x_2)\mathrm{log}_2(1x_2)(1+x_2)\mathrm{log}_2(1+x_2)],$$ (24) where $`x_{1,2}`$ can be written with the only parameter $`q=|R_{12}Q_{12}|`$ of the matrix $`R_{kl}Q_{kl}`$, diagonal matrix element $`p=\rho _{11}^A`$, and the coefficient module of correlations $`\mu =|\rho _{12}^A|/\sqrt{p(1p)}`$: $$x_1=\sqrt{14p(1p)(1q^2)},x_2=\sqrt{14p(1p)(1q^2\mu ^2)}].$$ The dependency corresponding to the Eq. (24) for $`p=1/2`$ (for the maximally possible amount of information of the source equal to 1 bit) is shown in Fig. 2a. ### V.2 Semiclassical information in the object–meter channel When one uses the object as a source of purely classical information in the most general form of the mixed ensemble $`\{p_\lambda ,\widehat{\rho }^A(\lambda )\}`$ the semiclassical information retrieved by the meter is described with the respective ensemble $`\{p_\lambda ,\widehat{\rho }_\lambda \}`$ resulted after the averaging over the object variables: $$\widehat{\rho }_\lambda =\underset{k}{}\rho _{kk}^A(\lambda )\left|k\right)\left(k\right|.$$ (25) Non-classicality of this channel is related to its nonzero commutator $`\widehat{C}_{\lambda \mu }=[\widehat{\rho }_\lambda ,\widehat{\rho }_\mu ]=_k[\rho _{kk}^A(\lambda )\rho _{ll}^A(\mu )\rho _{kk}^A(\mu )\rho _{ll}^A(\lambda )]Q_{kl}\left|k\right)\left(l\right|`$, which is nonzero only for the soft measurements with $`Q_{kl}\delta _{kl}`$. Nonorthogonality of the measurable states $`\left|k\right)`$ in ensemble (25) leads to the respective reduction of the retrieved information. Let us illustrate how the amount of information depends on the parameters of the soft measurement for the case of the multiple measurements in the same basis for the input ensemble of pure states $`\widehat{\rho }_k^A=|kk|`$. As one can easily see, this ensemble corresponds to another one of the form pure states $$\widehat{\rho }_k^{B^n}=||kk||$$ in the active subspace of the meter, which is described in Sec. IV.2. As an adequate quantitative characteristic for this kind of the channels we can use the semiclassical information $`I_s=S[p_k\widehat{\rho }_k^{B^n}]p_kS[\widehat{\rho }_k^{B^n}]`$ Holevo , which in this case is simply equal to the entropy $`S[\widehat{\rho }^{B^n}]`$ of the resulted density matrix $`\widehat{\rho }^{B^n}=p_k||kk||`$, which have the matrix elements (13) with $`\rho _{kk}^Ap_k`$, at the output of the measurement channel. For the two-level case, on account of Eq. (19) and for the a priori distribution $`p_k=\{1/2,1/2\}`$ we have the following amount of information: $$I_s=\frac{1}{2}\left(\mathrm{log}_2\frac{1e^{4\varkappa t}}{4}+e^{2\varkappa t}\mathrm{log}_2\frac{1+e^{2\varkappa t}}{1e^{2\varkappa t}}\right),$$ which monotonously changes from zero up to $`I_{\mathrm{max}}=1`$ with the change of the measurement time $`0\varkappa t\mathrm{}`$ (in arbitrary units). Effects of the external dephasing (reservoir) are not important in this case because related phases of the measurable states of the meter are not essential. ## VI Competition at the retrieval of the information from the object ### VI.1 Competition at the retrieval of the coherent information Let us consider now restrictions on the attained in the quantum measurement information for the case when there is a single source of information (we will call it “Alice”) and two receivers (“Bob” and “Eve”), which retrieve this information in series with the help of the repeated nondemolition quantum measurement with a different choice of the measurable variables, in general case (by contrast with Sec. IV.2). The mapping of the quantum states in the respective Hilbert spaces has the form $`H_AH_AH_EH_B`$ and the respective complete superoperator transformation can be written as $$_{EB}=\left(_E_B\right)\left(_E_B\right),$$ (26) where $$_E=R_{k_El_E}^E|k_E\left|k_E\right)\left(l_E\right|l_E|k_E||l_E,_B=R_{k_Bl_B}^B|k_B\left|k_B\right)\left(l_B\right|l_B|k_B||l_B$$ describe the measurements performed by Bob and Eve under the same object, but using different meters, and $`_{E,B}`$ is the respective identical transformation over the variable of the meter inaccessible in this measurement. Expanding Eq. (26), we have $$_{EB}=R_{k_El_E}^ER_{k_Bl_B}^Bk_B|k_El_E|l_B|k_B\left|k_E\right)\left|k_B\right)\left(l_B\right|\left(l_E\right|l_B|k_E||l_E.$$ In the specific case of coinciding meter’s bases , $`|k_B=|k_E`$, we have $$_{EB}^0=R_{kl}^ER_{kl}^B|k\left|k\right)\left|k\right)\left(l\right|\left(l\right|l|k||l.$$ The difference between the considered above case and that one described in Subsection IV.2 is in the independent use of information contained here in bipartite states $`\left|k\right)\left|k\right)`$ of the Eve–Bob system, which are coherently connected with the states $`|k`$ of the object. The result of the Eve’s measurement does not depend on the subsequent Bob’s measurement only for the marginal states $`\widehat{\rho }^B`$, but not for the joint states $`\widehat{\rho }^{AE}`$, which after the Bob’s measurement (for instance in the same basis) are represented instead of the initial superoperator $`_E=R_{kl}^E|k\left|k\right)\left(l\right|l|k||l`$ with the superoperator $$_{E}^{}{}_{}{}^{}=\mathrm{Tr}_B_{EB}^0=R_{kl}^ER_{kl}^BQ_{kl}^B|k\left|k\right)\left(l\right|l|k||l,$$ i.e. contains an additional dephasing factor $`R_{kl}^BQ_{kl}^B`$. An absence of the back action, i.e., the equality $`_{E}^{}{}_{}{}^{}=_E`$, in the case of $`R_{kl}^E1`$ when Eve retrieves the information in essentially quantum form, is occurred only for the completely coherent measurement by Bob, which contains no resulting information ($`R_{kl}^B1`$, $`Q_{kl}^B1`$). In the case of completely dequantized measurement by Eve there is never any reaction after the Bob’s measurement, which ensures the stability of classical information against its copying. A similar action does the measurement by Eve on the Bob’s measurement, which has after the Eve’s measurement the form $`_{B}^{}{}_{}{}^{}=\mathrm{Tr}_E_{EB}^0=R_{kl}^BR_{kl}^EQ_{kl}^E|k\left|k\right)\left(l\right|l|k||l,`$ i.e., contains an additional factor $`R_{kl}^EQ_{kl}^E`$ comparing to the case without Eve’s measurement. Such reaction of the quantum operation can be adequately described with the respected reduction of the coherent information due to its reception by a new receiver. In this case, due to the quantum measurement the meter receives the coherent information about the object states only after the measurement and the received information can be considered as the corresponding degree of quantum entanglement in the object–meter system, which is measured, for instance, in the system $`A+B`$ with the help of the difference $`I_c=S[\mathrm{Tr}_B\widehat{\rho }_A]S[\widehat{\rho }_A]`$, which is always positive in case of the soft measurement. By contrast with the similar definition used in ref. PRA2003, for the special case of a distinct entangling measurement, the first term here determines the entropy of the object, but not the meter, because the entropies of the object and the meter do not coincide for the case of the soft measurement. Calculating the respected information for Eve and Bob, we have $$\begin{array}{c}I_c^E=S\left[\left(\rho _{kl}^AR_{kl}^ER_{kl}^BQ_{kl}^EQ_{kl}^B\right)\right]S\left[\left(\rho _{kl}^AR_{kl}^ER_{kl}^BQ_{kl}^B\right)\right],\\ I_c^B=S\left[\left(\rho _{kl}^AR_{kl}^ER_{kl}^BQ_{kl}^EQ_{kl}^B\right)\right]S\left[\left(\rho _{kl}^AR_{kl}^BR_{kl}^EQ_{kl}^E\right)\right].\end{array}$$ The coherent information retrieved by Eve after Bob’s measurement and the amount of information received by Bob are shown for the two-dimensional case in Fig. 2b as functions of the softness parameters for the Bob’s ($`q_B=|Q_{12}^B|`$) and the Eve’s ($`q_E=|Q_{12}^E|`$ ) measurements at $`R_{12}^B=R_{12}^E=1`$ and for the density matrix of the object $`\rho _{11}^A=\rho _{22}^A=1/2`$, $`\rho _{12}^A=\rho _{21}^A=\mu /2`$. The respective analytical expressions for $`I_c^E,I_c^B`$ can be obtained using Eq. (24) for $`I_c(q,\mu )`$ that determines the coherent information about the reference system, which is preserved in the object with the initial density matrix $`\widehat{\rho }^A`$ after the soft measurement with the softness parameter $`q`$. This case corresponds to the change of variables $`\{qq_B\mu ,\mu q_E\}`$ when calculating information retrieved by Eve and, respectively, $`\{qq_E\mu ,\mu q_B\}`$ for the calculation of information retrieved by Bob. In these calculations, $`q_B=1`$ corresponds to the case when Bob practically does not perform the measurement and retrieves the unperturbed amount of Eve’s information $`I_c^E`$ (and vice versa), which is decreased with decreasing $`q_B`$ due to the competence. Note that the impact of the coherency parameter of the initial state $`\mu `$ on the coherent information (24), shown in Fig. 2a, and the competitive information $`I_c^{E,A}`$, shown in Fig. 2b, is opposite. Whereas the value of (24) with increasing $`\mu `$ falls due to decreasing of the initial entropy of the density matrix, which determines the entanglement between the object and the reference system, the information $`I_c^{E,A}`$ increases with increasing $`\mu `$ due to the respective increase of the object–meter entanglement after the measurement. This entanglement does not exist for the incoherent mixture of pure states $`\{p_k,|k\}`$ because of their imperturbability the states $`|k`$ even at the completely coherent measurement are described with the incoherent mixture of independent states $`|k\left|k\right)`$ of the object–meter system. The competition character for the selection of the coherent information reveals in an opposite action of the parameters $`q_B`$, $`q_E`$ on the information $`I_c^E`$ retrieved by Eve, for instance: with decreasing $`q_E`$, i.e., with increasing accuracy of the Eve’s measurement, her information increases, whereas with decreasing $`q_B`$ it is also decreases up to the zero at $`q_B=0`$ due to the partial transfer of the information to Bob. ### VI.2 Competition for selection of the classical information The competitive character of the quantum information is revealed also in the case of semiclassical channels $`AE`$, $`AB`$ with the given ensemble of input states. Ensemble corresponding to the first channel (25), as it can be easily checked by respective averaging of the superoperator (26), is not modified after the secondary measurement by Bob because he does not affects the input state of the object. However, the measurement result by Bob depends on the basis choice, which is used for the Eve’s measurement. The respective interdependency of the resulting quantum transformations lies in the basement of the quantum cryptography QC . Really, whereas the transformation $`_{AE}=_k\left|k\right)\left(k\right|k||k`$ includes only parameters of the measurement performed by Eve, namely, the measurable states $`|k`$ and the states of the meter $`\left|k\right)`$, the transformation of the channel $`AB`$ depends also on the parameters Eve’s measurement, in accordance with (26), as $$_{AB}=\underset{kl}{}R_{kl}^EQ_{kl}^E\left(𝒫_B|k_El|_E\right)k|_E|l_E.$$ (27) Here $`𝒫_B=_k\left|k\right)_B\left(k\right|_Bk|_B|k_B`$ is the superoperator of the *\[fuzzy\] soft projection* from $`H_A`$ onto $`H_B`$, which describes the result of the secondary measurement performed by Bob after the Eve’s measurement; $`|k,l_{E,B}`$ are the vectors of the measurable basis states in the space $`H_A`$ of the object states for the measurements by Eve and Bob. Introduced by Bob incoherency of the measurement due to the averaging over the states of the object does not affect the retrieved by Bob information and the soft character of the measurement is described with the superoperator $`𝒫_B`$. At the same time, the incoherency of the Eve’s measurement and its soft character are reflected with the common matrix of dephasing (external in respect to Bob) $`q_{kl}=R_{kl}^EQ_{kl}^E`$, which describes the resulting degree of the “softness” of Eve’s measurement. When the bases coincide, i.e., $`|k_E=|k_B=|k`$, the Eve’s measurement does not affect the information retrieved by Bob because $`𝒫_B|k_El|_E=\delta _{kl}\left|k\right)_B\left(k\right|_B`$ and dependence on the parameters $`R_{kl}^E,Q_{kl}^E`$ is vanished. In this case, both Eve and Bob use only classically compatible information about the object, which surely lacks the specific quantum nature of the competition in its selection, and this information can be copied independently. Dependency of the resulting channel $`AB`$ on the Eve’s transformations is related exclusively on the lack of coincidence of their measurements bases, which makes essential quantum disturbance of the object state introduced by Eve at $`q_{kl}1`$. The corresponding generalization of the semiclassical channel (25) on account of its modification (27) of the respective transformation of its quantum input, has the form: $$\widehat{\rho }_\lambda =\underset{kl}{}\rho _{kl}^{A|E}(\lambda )q_{kl}𝒫_B|k_El|_E,$$ (28) where $`\rho _{kl}^{A|E}`$ is the density matrix of the object in the basis Eve performs the measurement. The respected dependency of the semiclassical information $`I_s`$ on the measurement parameters and on the input ensemble is shown in Fig. 3. At its maximum degree the competition of the measurements reveals at the orientation angle $`\vartheta =\pi /2`$, which, in the case of physical realization of the Hilbert space of the object as the polarization degree of the freedom of a photon, corresponds to the rotation of the linear polarization of the photon at 45. ## VII Conclusions In conclusion, we introduced the model for selection of quantum information with the help of the generalized quantum nondemolition measurement, which takes into account the entanglement effects and, in most explicit form, summarizes the fundamental differences between the quantum and classical information. In particular, the possibility to consider the controlled degree of Eve’s interference while she uses the fuzzy measurement in order to attain the secret information transmitted over the channel Alice–Bob, allows one using this model of quantum measurement as the most simple, but conceptual enough model for the interaction of the streams of quantum information in quantum cryptography. ###### Acknowledgements. This work was partially supported by RFBR grant No.04–02–17554.
warning/0506/nucl-th0506032.html
ar5iv
text
# A NEW APPLICATION OF THE GÜRSEY AND RADICATI MASS FORMULA ## 1 Introduction Different versions of Constituent Quark Models (CQM) have been proposed in the last decades in order to describe the baryon properties. What they have in common is the fact that the three-quark interaction can be separated in two parts: the first one, containing the confinement interaction, is spin and flavour independent and it is therefore $`SU(6)`$ invariant, while the second violates the $`SU(6)`$ symmetry. This separation has been supported by the very first Lattice QCD calculations deru and is confirmed by the most recent ones bali12 ; alexandrou . The various CQMs differ in the way the $`SU(6)`$ invariance is violated. One of the most popular ways was the introduction of a hyperfine (spin-spin) interaction ik ; gi ; pl ; ff , however in many studies a spin- and isospin- bil ; olof ; ple ; iso or a spin- and flavour-dependent interaction bil ; olof ; ple has been considered. In this paper we study the symmetries of the baryon spectrum using a very simple approach based on the Gürsey Radicati mass formula (GR) GR . It is well known that the baryon spectrum exhibits an approximated $`SU(6)`$ symmetry and that the GR mass formula, despite it’s simplicity, describes quite well the way this symmetry is broken, at least in the lower part of the baryon spectrum. Our idea is to build up a very simple model based on the GR mass formula, to fix the parameters of the model in way to obtain the best description (within the limits of this approach) of the baryon spectrum and thereafter use the model (with the parameters values fixed on the baryon spectrum) to give predictions for the masses of other hadronic systems as for example pentaquarks bgs . The model we propose is a simple CQM where the $`SU(6)`$ invariant part of the Hamiltonian is the same as in the Hypercentral Constituent Quark Model pl ; ff and where the $`SU(6)`$ symmetry is broken by a GR inspired interaction. In the second section we briefly remind the hypercentral CQM and its results, then in the third section we construct the model using the GR as the $`SU(6)`$ breaking term. In the fourth section we give the results obtained by fitting the GR parameters to the strange and non strange baryon energies and we compare the spectrum with the experimental data. Finally we discuss our results. ## 2 The hypercentral model The experimental $`4`$ and $`3`$star non strange resonances can be arranged in $`SU(6)`$ multiplets. This means that the quark dynamics has a dominant $`SU(6)`$ invariant part, which accounts for the average multiplet energies. In the Hypercentral Constituent Quark Model (hCQM) it is assumed to be given by the hypercentral potential pl $$V(x)=\frac{\tau }{x}+\alpha x,$$ (1) where $$x=\sqrt{𝝆^2+𝝀^2},$$ (2) is the hyperradius and $`𝝆`$ and $`𝝀`$ are the Jacobi coordinates describing the internal quark motion. Interactions of the type linear plus Coulomb-like have been used since long time for the meson sector, e.g. the Cornell potential. This form has also been obtained in recent Lattice QCD calculations bali12 ; alexandrou for $`SU(3)`$ invariant static quark sources. Introducing, along with the hypercentral potential, a standard hyperfine interaction ik which breaks the $`SU(6)`$ symmetry, the hCQM has given a fair description of the non strange baryon spectrum pl and of other various physical quantities of interest: the photocouplings aie , the electromagnetic transition amplitudes aie2 , the elastic nucleon form factors mds and the ratio between the electric and magnetic proton form factors rap . Subsequently, in order to improve the description of the non strange spectrum, an isospin dependent $`SU(6)`$ violating term has been introduced iso . The complete interaction used in this latter case is given by $$H_{int}=V(x)+H_\mathrm{S}+H_\mathrm{I}+H_{\mathrm{SI}},$$ (3) where $`V(x)`$ is the linear plus hypercoulomb SU(6)-invariant potential of eq.(1), while $`H_\mathrm{S}`$ is the hyperfine interaction and $`H_\mathrm{I}`$, $`H_{\mathrm{SI}}`$ are respectively isospin and spin-isospin dependent terms. Similar results can be obtained in a relativized version of the model rel , in which the quark kinetic energy has the correct relativistic form. The preceding results show that both spin and isospin (or flavour) dependent terms in the quark Hamiltonian are important. Their contributions can be considered as perturbative $`SU(6)`$-violating terms added to the unperturbed $`SU(6)`$-invariant energies provided by the hypercentral potential of eq. (1). ## 3 The strange baryon spectrum and the Gürsey Radicati mass formula The spin and isospin dependent interactions considered in the previous section are not the only source of $`SU(6)`$ violation. In order to study the strange baryon spectrum one has to consider the $`SU(3)`$ violation produced by the differences in the quark masses. The well known Gell-Mann-Okubo (GMO) mass formula gmo made use of a $`\lambda _8`$ violation of $`SU(3)`$ in order to describe the mass splittings within the various $`SU(3)`$ multiplets; according to this formula the mass $`M`$ of a baryon belonging to a given $`SU(3)`$ multiplet can be expressed as $$M=M_0+DY+E\left[T(T+1)\frac{1}{4}Y^2\right],$$ (4) where $`M_0`$ is the average energy value of the $`SU(3)`$ multiplet, $`Y`$ is the hypercharge, $`T`$ is the Isospin of the baryon and $`D`$ and $`E`$ are parameters to be fitted. A simple way to interpret the origin of such a violation is just to attribute to the strange quark a mass different from the up and down quark ones. The calculations were performed without reference to any explicit dynamical model, but using standard group theoretical methods. The unknown parameters $`D`$ and $`E`$ in the $`SU(3)`$ violating terms can be in principle fitted to the experimental masses, thereby providing a phenomenological way to describe the spectrum. A similar approach for the description of the splittings within the SU(6) baryon multiplets is provided by the Gürsey Radicati mass formula GR . In the original paper the mass formula reads: $$M=M_0+CS(S+1)+DY+E\left[T(T+1)\frac{1}{4}Y^2\right]$$ (5) where $`S`$ is the spin. Eq.(5) can be rewritten in terms of Casimir operators in the following way $`M=M_0+CC_2[SU_S(2)]+DC_1[U_Y(1)]`$ $`+E\left[C_2[SU_I(2)]{\displaystyle \frac{1}{4}}(C_1[U_Y(1)])^2\right]`$ (6) where $`C_2[SU_S(2)]`$ and $`C_2[SU_I(2)]`$ are the $`SU(2)`$ (quadratic) Casimir operators for spin and isospin, respectively, and $`C_1[U_Y(1)]`$ is the Casimir for the $`U(1)`$ subgroup generated by the hypercharge $`Y`$. The presence of a spin dependent term is necessary since states belonging to a definite $`SU(6)`$ multiplet do not have the same spin value. This mass formula has proven to be successful in the description of the gruond state baryon masses, however, as stated by the authors themselves, eq. (3) is not the most general mass formula that can be written on the basis of a broken $`SU(6)`$ symmetry. In order to generalize eq. (3) one can consider a dynamical spin-flavour symmetry $`SU_{SF}(6)`$ and write the following chain of subgroups $$\begin{array}{cccccccccccc}SU_{SF}(6)& & SU_F(3)& & SU_S(2)& & SU_I(2)& & U_Y(1)& & SO_S(2)& \\ & & & & & & & & & & & \\ (\lambda _1,..\lambda _5)& & (\lambda _f,\mu _f)& & S& & I& & Y& & M_S& \end{array}$$ (7) where in the bottom row we report the quantum numbers which label the irreducible representations of the corresponding groups. Therefore one can describe the $`SU_{SF}(6)`$ symmetry breaking mechanism by generalizing eq. (3) as $`M=M_0+AC_2[SU_{SF}(6)]+BC_2[SU_F(3)]`$ $`+CC_2[SU_S(2)]+DC_1[U_Y(1)]`$ $`+E\left(C_2[SU_I(2)]{\displaystyle \frac{1}{4}}(C_1[U_Y(1)])^2\right)`$ (8) The generalized Gürsey Radicati mass formula eq. (3) can be used to describe the whole baryon spectrum, provided that two conditions are fulfilled. The first condition is the possibility of using the same splitting coefficients for different $`SU(6)`$ multiplets. This seems actually to be the case, as shown by the algebraic approach to the baryon spectrum bil , where a formula similar to eq. (3) has been applied. The second condition is given by the possibility of getting reliable values for the unperturbed mass values $`M_0`$. Our idea is to use for this purpose the $`SU(6)`$ invariant part of the hCQM, which provides a good description of the non strange baryon spectrum and to introduce a Gürsey Radicati inspired $`SU(6)`$ breaking interaction to describe the splittings within each $`SU(6)`$ multiplet. We shall therefore make use of the following three quark Hamiltonian $$H=H_0+H_{GR}$$ (9) with $$H_0=3m+\frac{𝒑_\lambda ^2}{2m}+\frac{𝒑_\rho ^2}{2m}+V(x),$$ and $$H_{GR}=+AC_2[SU_{SF}(6)]+BC_2[SU_F(3)]+CC_2[SU_S(2)]$$ $$+DC_1[U_Y(1)]+E\left(C_2[SU_I(2)]\frac{1}{4}(C_1[U_Y(1)])^2\right),$$ where $`V(x)`$ is the hypercentral potential of eq.(1), and m is the constituent quark mass. It has to be remarked that, in order to simplify the solving procedure, the constituent quark masses are assumed to be the same for all the quark flavours ($`m_u=m_d=m_s=m`$), therefore, within this approximation, the $`SU(3)`$ symmetry is only broken dynamically by the spin and flavour dependent terms in the Hamiltonian. In other words, in this approximation, the effects of the strange quark mass on the baryon spectrum are described by the two terms of eq.(4). The eigenproblem of $`H_0`$ can be solved numerically, the spin-flavour part of the resulting eigenfunctions has definite properties under transformations of the $`SU_{SF}(6)`$ group and its subgroups. Using the notation of equation (7) the spin-flavour part of the wave function can be written as $$|(\lambda _1,\lambda _2,\lambda _3,\lambda _4,\lambda _5),(\lambda _f,\mu _f),I,Y,S,M_S.$$ (10) Often the irreducible representations are identified not by the quantum numbers but by their dimension. Thus, for example $$|(3,0,0,0,0),(2,1),I=\frac{1}{2},Y=1,S=\frac{1}{2},M_S$$ $$|^28_{1/2},[56,0^+],N$$ is the spin-flavour part of the nucleon wave function. The notation used is $$|^{2S+1}\text{dim}(SU(3))_J,[\text{dim}(SU(6)),L^P],X,$$ where dim$`(SU(n))`$ is the dimension of the $`SU(n)`$ representation, $`S`$ and $`L`$ are the total spin and orbital angular momentum of the quark system, respectively, $`J`$ and $`P`$ are the spin and parity of the resonance and $`X=N,\mathrm{\Delta },\mathrm{}`$ denotes the type of baryon resonance. The action of $`H_{GR}`$ on the eigenstates of $`H_0`$ is completely identified by the expectation values of the Casimir operators on the states of eq. (10) $`C_2[SU_{SF}(6)]`$ $`=`$ $`\{\begin{array}{ccc}45/4& \text{for}& [56]\\ 33/4& \text{for}& [70]\\ 21/4& \text{for}& [20]\end{array}`$ (14) $`C_2[SU_F(3)]`$ $`=`$ $`\{\begin{array}{ccc}3& \text{for}& [8]\\ 6& \text{for}& [10]\\ 0& \text{for}& [1]\end{array}`$ (18) $`C_2[SU_I(2)]`$ $`=`$ $`I(I+1)`$ $`C_1[U_Y(1)]`$ $`=`$ $`Y`$ $`C_2[SU_S(2)]`$ $`=`$ $`S(S+1)`$ (19) Therefore the mass of each baryon state $`|B`$ can be written as: $$B|H|B=E_{\gamma \nu }+B|H_{GR}|B$$ (20) where $`E_{\gamma \nu }`$ denotes the eigenvalue of $`H_0`$ with $`\gamma =2n+l_\rho +l_\lambda `$ ($`n`$ being a non negative integer), $`\nu `$ denotes the number of nodes of the space three quark wave functions and $`l_\rho `$, $`l_\lambda `$ are the orbital angular momenta corresponding to the Jacobi coordinates (see e.g. pl ). Since $`H_{G.R.}`$ does not depend on the spatial degrees of freedom, the $`SU(6)`$ breaking term introduced in this model is diagonal in the baryon states, this means that the Gürsey Radicati term is able to give energy splittings within the $`SU(6)`$ multiplets, but no configuration mixing effects can arise from such an interaction <sup>1</sup><sup>1</sup>1The kind of problems that one can face neglecting the spatial dependence on the $`SU(6)`$ breaking part is discussed by Jennings and Maltman Jennings:2003wz . Therefore the model is expected to fail in the description of all thoose observables where a good description of the three quark wave function is crucial. ## 4 Results The parameters in $`H_{GR}`$ can be determined in order to reproduce the experimental values of the energy splittings. We first adopt an analytical procedure by means of which we choose a limited number of well known resonances and express their mass differences using $`H_{GR}`$ and the Casimir operator expectation values given in the previous section. We list in the following the analytical expressions for the mass differences of the chosen pairs of resonances: $`(N(1650)S11N(1535)S11)`$ $`=`$ $`3C`$ $`(\mathrm{\Delta }(1232)P33N(938)P11)`$ $`=`$ $`9B+3C+3E`$ $`(N(1535)S11N(1440)P11)`$ $`=`$ $`(E_{10}E_{01})+12A`$ $`(\mathrm{\Sigma }(1193)P11N(938)P11)`$ $`=`$ $`{\displaystyle \frac{3}{2}}ED`$ $`(\mathrm{\Lambda }(1116)P01N(938)P11)`$ $`=`$ $`D{\displaystyle \frac{1}{2}}E.`$ (21) Looking at eq. (4) it is easy to understand that for the description of the non-strange baryon spectrum the only relevant parameters are $`A`$, $`C`$ and the combination $`(3B+E)`$. It should be noted that, in order to apply the Gürsey Radicati mass formula to the excited states, it is necessary to know the coefficient $`A`$ of the $`SU(6)`$ Casimir operator and the excited energies provided by the CQM. Once the SU(6) breaking interaction has been determined, the parameters of $`H_0`$ ($`\alpha `$ and $`\tau `$) which lead to the unperturbed energies $`E_{\gamma \nu }`$ can be fixed by a minimization procedure on the non-strange baryon spectrum. The complete list of the parameter values is reported in Table 1, column (I). In this way the $`E_{\gamma \nu }`$ levels are placed close to the central mass value of each SU(6) multiplet. As shown in eq. (4), a further adjustment to the unperturbed multiplet energy is provided by the presence of the $`SU(6)`$ Casimir operator. The resulting spectrum is shown in Fig. 1 and Table 2 column $`M_{calc}^I`$. Despite of the simple form of the SU(6) breaking interaction, the general features of the spectrum are fairly well reproduced, especially in the low energy part. It has to be noted in particular that, besides the ground state masses which have been fixed through eq.(4), the predicted masses of the $`\mathrm{\Sigma }^{}`$, $`\mathrm{\Xi }`$, $`\mathrm{\Xi }^{}`$ and $`\mathrm{\Omega }`$ states are nicely close to the experimental values. The second approach followed in the application of the Gürsey Radicati mass formula is to fit all parameters at the same time in order to obtain the best reproduction of the spectrum of the 3 and 4 star resonances <sup>2</sup><sup>2</sup>2The PDG pdg quotes a three stars $`\mathrm{\Xi }(1690)`$ resonance; however, since the values of spin and parity are not known, this resonance cannot be identified with a definite eigenstate of the Hamiltonian and therefore this state has been excluded from our analysis.. The fitted parameters are reported in Table 1, column (II), while the resulting spectrum is shown in Fig. (2) and the corresponding numerical values are given in Table 2, column $`M_{calc}^{II}`$. The result is a better overall agreement with the experimental data, even if the prediction in the non strange sector is worsened. For this reason, we prefer the values of the parameters obtained with the previous analytical method since we have used only well known and well established resonances in order to fix the parameters. In both cases a non zero value of the $`SU(6)`$ Casimir coefficient is needed in order to reproduce the average multiplet energies. We have also tried a fit imposing $`A=0`$. The resulting parameters $`\alpha `$ and $`\tau `$ are however considerably different with respect to those of Table 2 because the lack of the parameter $`A`$ must be compensated by the $`SU(6)`$ invariant energies provided by the hypercentral potential. This is particularly evident in the case of the negative parity resonances, where the energy difference $`E_{10}E_{01}`$ must be bigger than in the previous case in order to obtain a good reproduction of the masses; in this way, however, the right ordering of the Roper resonance and the negative parity resonances is lost. This means that the presence of the Casimir $`C_2[SU_{SF}(6)]`$ is needed and its effect is to shift down the energy of the first excited $`0^+`$ state with respect to that of the $`1^{}`$. The effect of this term is very similar to that of the phenomenological $`U`$ potential of the Isgur-Karl model ik The mass formula of eq.(3) can be used to add a simple SU(6) breaking interaction to a CQM and despite its simplicity it gives rise to a good description of the baryon spectrum. Of course for the wavefunctions, and other observables, it is not expected to be as successful as for the spectrum. Another important feature of this kind of approach is the model independence of the $`SU(6)`$ breaking part of the Hamiltonian. Looking at eq.(4) it is easy to understand that the values of the parameters of the SU(6) breaking part of the Hamiltonian (i.e. the B,C,D,E parameters) are completely independent on the choice of $`H_0+AC_2[SU_{SF}(6)]`$ which must describe the central mass value of each $`SU(6)`$ multiplet. An important consequence of the independence, in first approximation, of the coefficients of the Casimir operators on the particular wave functions, is the possibility of using the same $`SU(6)`$ breaking Hamiltonian for systems with a different number of quarks. This has been done in a recent study of the pentaquark spectrum, where the Gürsey Radicati mass formula of eq.(3) has been used for a systematic analysis of the ($`SU(6)`$ breaking) splittings within the exotic baryon multiplets bgs . Finally, we present some comments on the Gürsey Radicati mass formula of eq.(3). As we have already observed, the last two terms, that is those with coefficients $`D`$ and $`E`$, describe up to first order the $`SU(6)`$-violation coming from the mass difference of quarks, as it has been done in the Gell-Mann-Okubo formula. The remaining terms are expected to appear once an explicit dynamics for the quark system is introduced. For example, in a recent calculation of multiquark state energies, a spin-flavour dependent interaction of the type $$H_{\lambda \sigma }=\underset{i<j}{\overset{n}{}}(𝝀_i𝝀_j)(𝝈_i𝝈_j)$$ has been introduced helminen ; $`\lambda _i`$ are the $`SU_{flavour}(3)`$ matrices. The matrix elements of such spin-flavour interaction between states belonging to definite irreducible representations of $`SU(6)`$, $`SU(3)_{flavour}`$ and $`SU(2)_{spin}`$ can be calculated as $`[f]^{SU(6)}[f]^{SU(3)}[f]^{SU(2)}\left|H_{\lambda \sigma }\right|[f]^{SU(6)}[f]^{SU(3)}[f]^{SU(2)}`$ $`=4C_2(SU(6))2C_2(SU(3)){\displaystyle \frac{4}{3}}C_2(SU(2))8N_q,`$ (22) where $`N_q`$ is the number of quarks helminen and $`C_2(SU(2))`$ is given by $`S(S+1)`$, $`S`$ being the total spin. If the spatial dependence of the $`SU(6)`$ breaking terms is not neglected, this is no more true. As a conclusion, we can say that the Gürsey Radicati (3) is a simple way to parametrize at the first order the possible $`SU(6)`$-breaking terms of the strong interaction. The approach we have adopted here is then to parametrize all the $`SU(6)`$-breaking terms by means of the generalized Gürsey Radicati, without formulating any hypothesis on the microscopic mechanism (one-gluon exchange, Goldstone-boson interaction, chiral soliton …). The next step will be to introduce an explicit $`SU(6)`$ breaking term, containing also the spatial dependence and with a clear microscopic origin. ## 5 Discussion We have shown that the Gürsey Radicati (GR) mass formula is a good parametrization of the baryon energy splittings coming from $`SU(6)`$ breaking. The splittings are considered as perturbations superimposed to the $`SU(6)`$ invariant levels, which, in our approach, are given by the hypercentral three-quark potential pl . The overall good description of the spectrum which we obtain shows that the GR mass formula can also be used to give a fair description of the energies of the excited multiplets at least up to 2 GeV and not only for the ground state octets and decuplets, where it has been originally applied. There are still problems with the reproduction of some hyperons, in particular for the $`\mathrm{\Lambda }(1405)`$ and the $`\mathrm{\Lambda }(1520)`$ resonances that come out degenerate and above the experimental values. There are problems in the reproduction of the experimental masses also in the $`\mathrm{\Sigma }`$ sector where both the $`\mathrm{\Sigma }(1670)D13`$ and the $`\mathrm{\Sigma }(1775)D15`$ four stars resonances turn out to have predicted masses about $`100`$ MeV above the experimental values. A better agreement can be obtained either using the square of the mass bil or trying to include a spatial dependence in the $`SU(6)`$-breaking part, which may have, among others, a delta or Gaussian factor, in order to decrease the breaking with the increase of the spatial excitation. Although the space dependence of the $`SU(6)`$ breaking terms has to be neglected in order to apply the GR formula, We can consider the Gürsey Radicati $`SU(6)`$ breaking as the first order parametrization of the splittings due to an interaction which depends also on the coordinates. Within this approximation it can be used for the description of the SU(6)-breaking effects independently from the way in which one describes the spatial part. It can be applied, with the same coefficients, also to systems with different number of quarks, such as baryons or pentaquarks. A similar statement is valid if we restrict ourselves to $`SU(3)`$ breaking, using a Gell-Mann-Okubo mass formula. In fact, the representations involved will differ for each system and the dependence on the number of quarks will be accounted for by the different values of the various Casimir operators. On the contrary, the unperturbed $`SU(6)`$ invariant levels will depend on the number of quarks and on the way in which the spatial part is described, that is an explicit dynamics must be considered. What we have presented here is not the only example of such a situation. In fact recently in diakonov a Hamiltonian, containing the quadratic $`SU(3)`$ Casimir and a Gell-Mann Okubo symmetry breaking term, has been used for the calculations of energy splittings both for baryons and pentaquarks. This is an indication that different effective models for the baryons at the first oder give origin to a Gell-Mann Okubo mass formula, independently from the fact that we consider a Chiral Soliton Model or a bag model or a CQM, that means independently from which effective degrees of freedom we use and how we describe from a spatial point of view the baryon bound states.
warning/0506/hep-th0506197.html
ar5iv
text
# 1 Introduction ## 1 Introduction String theory should provide us with the dynamics and geometry of target space. In fact it should allow us to understand precisely what target space is and to what extent the classical concepts of space and time are valid. The discovery of branes and dualities in string theory has made this a vast topic. Noncritical string theories with $`c1`$ defined via matrix models allow us to study various nonperturbative aspects of string theory. In particular they offer us a possibility to study the summation of string perturbation theory, which was why they attracted so much attention in the early 90ties. They play a role similar to the one that certain 2d quantum field theories play in the study of 4d quantum field theories: the possibility to solve them allows us to test nonperturbative concepts. More recently the study of noncritical strings has again received much attention since they have a worldsheet formulation as conformal field theories coupled to Liouville field and an explicit matrix model formulation. Important progress has been made in both formulations in the understanding of $`D`$-branes and has resulted in a fruitful interplay between the two ”dual” theories. In particular for $`c<1`$, where a target space formulation had been missing, the study of $`D`$-branes gave us an understanding of the target space via the study of the moduli space of the D-branes. The space of the so-called FZZT-branes is labeled by a single parameter $`x`$ and the disk amplitude $`S(x)`$ satisfies the following (semiclassical) equation $$T_l(W/C)=T_k(x),W=_xS(x)$$ (1) in the $`(x,W)`$ plane, where $`C`$ is a constant and $`T_p`$ are Chebyshev polynomials. Eq. (1) is valid for a minimal $`(l,k)`$ string theory in the so-called conformal closed string background. For future reference we note that in the special case of a minimal $`(2,2m1)`$ string theory, eq. (1) can be written as $$W^2=\left(P_m(x,u)\right)^2(x+u),$$ (2) where $`P(x,u)`$ is a polynomial of degree $`m1`$ in $`x`$ and $`u`$. $`u=\sqrt{\mu }`$ where $`\mu `$ can be viewed as an effective cosmological constant and $`x`$ can be interpreted as a boundary cosmological constant. We also note that (2) is valid for any so-called closed string background, $`u`$ being a function $`u(t)`$ of a certain number of coupling constants $`t_i`$, $`i=0,\mathrm{},m`$, as will be explained later. From this point of view the semiclassical moduli space is a rather complicated Riemann surface (2) and in the ordinary complex plane $`x`$ the function $`S(x)`$ has cuts. This picture is true to all finite orders in string perturbation theory. In fact it is essential for the whole philosophy of string perturbation theory in the context of matrix models that the cut structure in the complex plane is unchanged. However, since it was first realized that the string perturbation theory for non-critical strings could be resummed it was also realized that the resummed solution to string theory (the non-perturbative partition function in the presence of the FZZT brane labeled by $`x`$) was an entire function in the complex $`x`$-plane . The cut had disappeared and it was seemingly an artifact of the perturbative expansion. This point was reemphasized in a recent paper in the context of eq. (1). The purpose of this note is to explore what kind of geometrical structure should replace the Riemann surface (2) on the (nonperturbative) quantum level. In particular one has to give a meaning to the $`W`$ variable on the nonperturbative level since it does not make sense to consider disk amplitudes then. Promoting $`W`$ to an independent coordinate is also natural from the point of view of the link between $`c<1`$ strings and B-model topological string on the Calabi-Yau manifold $$zw=W^2P^2(x)(x+u)$$ (3) where $`W`$ and $`x`$ (and $`z`$, $`w`$) are (complex) coordinates. In this paper, however, we will not address this issue and stay purely within the context of $`c<1`$ strings. The ‘quantum’ $`W`$ and $`x`$ coordinates are in a completely natural way noncommutative while the vestige of the Riemann surface (2) on the quantum level is an additional projection operator. In this note we make the above explicit – we point out that the disappearance of the Riemann surface (1) has a natural interpretation in terms of noncommutative geometry: For a non-zero string coupling one can define the concept of noncommutative phase space and in some sense, to be specified below, the Riemann surface (1) appears in the limit of vanishing string coupling which is also the limit where the noncommutative parameter goes to zero. The formulation also has the advantage that it brings the variables $`x`$ and $`W`$ in (1) on an equal footing as phase space variables, it describes in a natural way the ”quantum mechanical” nature of target space and it connects to a similar (well known) description in the case of the $`c=1`$ string . The integrable KdV structure of non-critical string theories is related to a formulation in terms of free fermions (as is the case for most integrable systems) and it is this fermionic structure which allow us in a natural way to introduce a noncommutative phase space. In the next section we will present the noncommutative geometric structure, and we then study the explicit realization in $`(2,2m1)`$ non-critical string theory. ## 2 The noncommutative geometry framework The $`(2,2m1)`$ models have a representation via one-matrix integrals. Before the continuum limit is obtained we can write the partition function as $$Z=𝑑M\text{e}^{\frac{1}{g_s}\mathrm{Tr}V(M)}$$ (4) for a suitable potential $`V(M)`$. This matrix model can be solved using orthogonal polynomials $`P_n(x)`$ satisfying $$𝑑xP_n(x)P_m(x)\text{e}^{\frac{1}{g_s}V(x)}=\delta _{nm}$$ (5) A free fermion description can be obtained by introducing fermions with wave functions $$\psi _n(x)=P_n(x)\text{e}^{\frac{1}{2g_s}V(x)},$$ (6) i.e. we introduce a second quantized free fermion field $$\mathrm{\Psi }(x)=\underset{n=1}{\overset{\mathrm{}}{}}a_n\psi _n(x),\{a_m,a_n^{}\}=\delta _{m,n}.$$ (7) The Fermi sea is defined by filling up the first $`N`$ levels, where $`N`$ is the size of the matrix, i.e. $$a_n|0=0\mathrm{for}nN,a_n^{}|0=0\mathrm{for}n<N.$$ (8) Expectation values of traces of the matrix $`M`$ can now be calculated nicely in the free fermionic theory, i.e. $$\mathrm{Tr}M^k=N|\mathrm{\Psi }^{}\widehat{x}^n\mathrm{\Psi }|N,$$ (9) where $`\widehat{x}`$ is the multiplication operator corresponding to $`x`$. The origin of the fermionic nature of $`\mathrm{\Psi }`$ is the Vandermonde determinant resulting from the integration over the angular part of the matrix variables in the integral (4). The wave function of the first unoccupied level is $`\psi _N(x)`$, which by the Heine formula for orthogonal polynomials can be written: $$\psi _N(x)=det(xM),$$ (10) where the expectation value is taken with respect to partition function $`Z`$ in (4). The continuum limit of the $`(2,2m1)`$ matrix model is obtained by taking a certain scaling limit when $`N\mathrm{}`$ and the variable $`x`$ is close to the cut of eigenvalues of the matrix model in the large $`N`$ limit. More explicitly one writes: $$xx_c+ax,\frac{n}{N}=1a^m(t_0\tau ),Na^{m+\frac{1}{2}}=\frac{1}{g_s}.$$ (11) With this notation $`\tau t_0`$ means $`nN`$, and as will be described below, $`t_0`$ has the interpretation as the coupling constant related to a perturbation with the most dominant primary operator of the $`(2,2m1)`$ conformal field theory coupled to gravity, while $`x`$ has the interpretation of a boundary cosmological constant. In this limit one obtains (see for instance for a review) $$\psi (x_c+ax)\psi (\tau ,x),$$ (12) where the double-scaled function $`\psi (\tau ,x)`$ satisfies $$Q\psi (\tau ,x)=x\psi (\tau ,x),Q=g_s^2\frac{\text{d}^2}{\text{d}\tau ^2}u(\tau ,g_s).$$ (13) Eq. (13) can be viewed as a Schrödinger-like equation with $`Q`$ playing the role of the Hamiltonian, $`u(\tau )`$ the role of the potential and $`x`$ the role of energy. It would allow for a standard WKB expansion in $`g_s`$ if it was not for the fact that the ”potential” $`u(\tau ,g_s)`$ was itself a function of $`g_s`$, determined from the so-called string equation $$[Q,P]=g_s,P\psi (\tau ,x)=g_s\frac{\text{d}}{\text{d}x}\psi (\tau ,x)$$ (14) where $`P`$ is a polynomial in $`d/d\tau `$ and $`u`$. As is seen $`Q`$ and $`P`$, represent in a way the noncommutative phase space variables $`q=x`$ and $`p=g_sd/dx`$. $`p`$ is then to be identified with the $`W`$ coordinate on the (nonperturbative) quantum level. However, from (13) another (conceptually simpler) representation of a noncommutative phase space will emerge which we now turn to describe. The important point for us at this stage is the notion of a Fermi sea, the ground state, in a theory of free fermions, above which we consider excitations. We thus have to project from wave functions all components below the Fermi level, which is explicitly done by the kernel $$K_{c<1}(x,y)=\underset{n=N}{\overset{\mathrm{}}{}}\psi _n(x)\psi _n(y)$$ (15) It can be written as $$K_{c<1}(x,y)=\delta (xy)\underset{n=0}{\overset{N1}{}}\psi _n(x)\psi _n(y)=\delta (xy)K_{rmm}(x,y)$$ (16) where $$K_{rmm}\frac{\psi _N(x)\psi _{N1}(y)\psi _{N1}(x)\psi _N(y)}{xy}$$ (17) is the standard random matrix kernel where an appropriate double scaling limit should be taken according to (11). Clearly a derivative with respect to $`\tau `$ is involved in the double scaling limit of (17) and using eq. (13) one obtains $$K_{c<1}(x,y;t_0)=_{t_0}^{\mathrm{}}\text{d}\tau \psi ^{}(\tau ,x)\psi (\tau ,y)$$ (18) The first unoccupied level is then exactly the Baker-Akhiezer function (or the 1-point function of the FZZT brane) as is clear from (10) and in the double scaling limit we can write $`\psi (t_0,x)`$. The above construction fits nicely into the general framework of non-commutative geometry. Let the starting point be the noncommutative plane. It can be defined as the algebra $`𝒜`$ generated by $`p`$ and $`q`$ operators acting on $`=L^2()`$. This Hilbert space in the context of $`c<1`$ strings is naturally identified with the space of wavefunctions (12) of the excitations of the random matrix model in the double scaling limit (we will discuss the Fermi sea in this context shortly). In order to pass to the picture of functions of two variables (these will be the $`piW`$ and $`qx`$ of the target space of the minimal string theory) with noncommutative multiplication one uses the Weyl map $$f(p,q)\widehat{f}=\frac{\text{d}k_1\text{d}k_2}{(2\pi )^2}\text{e}^{i(k_1\widehat{q}+k_2\widehat{p})}\stackrel{~}{f}(k_1,k_2),$$ (19) where $`\stackrel{~}{f}`$ is the Fourier transform of $`f`$. Eq. (19) maps functions to operators, and defines the noncommutative multiplication by $$\widehat{f}\widehat{g}fg.$$ (20) The inverse map, sometimes called the Wigner map, is constructed by introducing via (19) the operator $`\widehat{\mathrm{\Delta }}`$ corresponding to the delta-function $`\delta (qq_0)\delta (pp_0)`$: $$\widehat{\mathrm{\Delta }}(q_0,p_0)=\frac{\text{d}k_1\text{d}k_2}{(2\pi )^2}\text{e}^{i(k_1\widehat{q}+k_2\widehat{p})}\text{e}^{i(q_0k_1+p_0k_2)},$$ (21) and it is $$\widehat{f}f(q,p)=\mathrm{Tr}\left(\widehat{f}\widehat{\mathrm{\Delta }}(q,p)\right)$$ (22) The noncommutative product defined in this way by (20) is exactly equivalent to the standard Moyal product: $$f(q,p)g(q,p)=f(q,p)\mathrm{exp}\left(i\frac{1}{2}\stackrel{}{_\mu }\theta _{\mu \nu }\stackrel{}{_\nu }\right)g(q,p)$$ (23) where $`\mu ,\nu `$ refers to the $`q,p`$ coordinates and $`\theta _{\mu \nu }=\theta \epsilon _{\mu \nu }`$ is the noncommutative parameter of the the theory. At this stage we see that one ingredient of the $`c<1`$ string theory has not yet been incorporated so far. Since the elementary excitations are fermions and all the levels in the Fermi sea are already filled, the excitations $`\psi (\tau ,x)`$ cannot probe the whole Hilbert space $`=L^2()`$ and one has to project out from $``$ the states in the Fermi sea through (15). This projection leads to an additional geometrical structure in the noncommutative plane which is essentially the quantum nonperturbative analog of the classical Riemann surface (2) as we will show. To this end let us note that associated with a given quantum wave function $`\psi (x)`$ we have the projection operator $`|\psi \psi |`$ and the Wigner map applied to this operator gives the Wigner function $`f_\psi (q,p)`$ associated with the wave function $`\psi (x)`$: $$f_\psi (q,p)=\frac{\text{d}x}{2\pi \mathrm{}}\text{e}^{ipx/\mathrm{}}\psi ^{}(qx/2)\psi (q+x/2).$$ (24) where $`q=(x_1+x_2)/2`$, the variables $`x_1`$ and $`x_1`$ being the coordinates which appear in the projector $`x_2|\psi \psi |x_1`$. Now we are ready to write the Wigner map associated with the projector $$\widehat{\rho }=\underset{n=N}{\overset{\mathrm{}}{}}|\psi _n\psi _n|$$ (25) which leads to the associated Wigner function $$f_\rho (q,p)=\frac{\text{d}x}{2\pi \mathrm{}}\text{e}^{ipx/\mathrm{}}\rho (qx/2,q+x/2),$$ (26) where again $`\rho (x_1,x_2)=x_1|\widehat{\rho }|x_2`$ and $`q=(x_1+x_2)/2`$. The projection acts then on functions of two variables $`q=x`$ and $`p=iW`$ by the star multiplication: $$f(q,p)f_\rho (q,p)f(q,p)$$ (27) The geometrical structure which appears in the examples considered in this paper supplements the noncommutative plane by an appropriate (orthogonal) projection operator $`\widehat{K}`$. One then truncates the space of functions to the image of $`\widehat{K}`$: $$_K=\{\widehat{K}f:fL^2()\}$$ (28) and similarly the algebra of operators acting on $`_K`$: $$𝒜_K\{\widehat{K}\widehat{O}:\widehat{O}𝒜\}=\{\widehat{K}\widehat{O}\widehat{K}:\widehat{K}𝒜\}$$ (29) It is seen that this structure is precisely the one present also in $`c=1`$ non-critical string theory . Let us now discuss in more detail the implementation of this abstract noncommutative geometrical structure in the $`(2,2m1)`$ minimal models. Before we proceed with the general discussion let us illustrate the above notions in the simplest example of $`(2,1)`$ strings. #### $`(2,1)`$ minimal string theory In this case the Riemann surface is $$W^2qt_0=0$$ (30) Let us now proceed to find the projector which represents the geometry of the $`(2,1)`$ theory according to (26). In this case the double scaling wavefunctions can be obtained exactly and we have $$\left(\mathrm{}^2\frac{d^2}{dt_0^2}t_0\right)\psi (t_0,x)=x\psi (t_0,x),\psi (t_0,x)=\mathrm{Ai}\left(\frac{x+t_0}{\mathrm{}^{2/3}}\right),$$ (31) while the Wigner transform is $$f_\psi (q,p)=\frac{1}{\pi \mathrm{}^{2/3}}\mathrm{Ai}\left(\frac{p^2+(q+t_0)}{\mathrm{}^{2/3}}\right)$$ (32) We see already here that the curve (30) appears as the argument of the Airy function. The Wigner transform of the projector is then obtained by integrating over $`t_0`$ $$f_\rho (q,p)=_{t_0}^{\mathrm{}}\text{d}t_0^{}f_\psi (q,p;t_0^{}),$$ (33) Let us now examine the classical limit $`\mathrm{}0`$. Using the known asymptotics of integrals of Airy functions we have in the classically forbidden region $$f_\rho (q,p)\frac{\mathrm{}^{1/2}}{(p^2+q+t_0)^{3/4}}\mathrm{exp}\left(\frac{2}{3\mathrm{}^{2/3}}(p^2+q+t_0)^{3/2}\right),$$ (34) while in the classically allowed region $$f_\rho (q,p)1\frac{\mathrm{}^{1/2}}{\sqrt{\pi }(p^2qt_0)^{3/4}}\mathrm{cos}\left(\frac{2}{3\mathrm{}^{2/3}}(p^2qt_0)^{3/2}+\frac{\pi }{4}\right).$$ (35) Therefore the projector in this limit approaches the Heaviside step function $$f_\rho (q,p)\theta (p^2+q+t_0)$$ (36) The Riemann surface curve appears then as the boundary of the classically allowed region which is natural since the FZZT 1-point function is the lowest excitation above the Fermi sea. We will now proceed to explore these issues in the much more complicated case of $`(2,2m1)`$ minimal string theories. First we outline how one obtains (1) in the semiclassical limit where $`\mathrm{}g_s0`$ from the Wigner-function corresponding to $`\psi (t_0,x)`$, viewed as a function of $`x`$, not $`t_0`$. ## 3 The semiclassical limit and beyond The semiclassical solution to $`(2,2m1)`$ non-critical strings is entirely determined by the disk-amplitude. It can be written as function of a number of coupling constants $`t_i`$, $`i=0,m`$, and the boundary cosmological constant $`q`$. For a given choice of $`t_i`$ the explicit formula for the semiclassical disk-amplitude is : $$S(q,t)=\frac{1}{2}_{\mathrm{}}\frac{\text{d}\mathrm{\Omega }}{2\pi i}V^{}(t,\mathrm{\Omega })\mathrm{log}\left(\frac{1+\sqrt{\frac{q+u(t)}{\mathrm{\Omega }+u(t)}}}{1\sqrt{\frac{q+u(t)}{\mathrm{\Omega }+u(t)}}}\right).$$ (37) while the formula for the derivative of the disk amplitude w.r.t. the boundary cosmological constant is (using the string equation (41)) $$W(q,t)=\frac{S}{q}=\frac{1}{2}\left(_{\mathrm{}}\frac{\text{d}\mathrm{\Omega }}{2\pi i}\frac{V^{}(t,\mathrm{\Omega })}{(\mathrm{\Omega }q)\sqrt{\mathrm{\Omega }+u(t)}}\right)\sqrt{q+u(t)},$$ (38) The (derivative of the) potential is defined by $$V^{}(t,q)=\underset{k=0}{\overset{m}{}}t_kq^{k1/2},$$ (39) and the function $`u(t)`$ is determined by the string equation (14). It can be written as a differential equation in $`u`$: $$\underset{k=0}{\overset{m}{}}t_kR_k(u)=0,$$ (40) where $`R_k(u)`$ denotes the so-called Gelfand-Dikii differential polynomials of $`u`$, considered as a function of $`t_0`$<sup>1</sup><sup>1</sup>1The Gelfand-Dikii differential polynomials can be recursively defined by $$\frac{\text{d}}{\text{d}t_0}R_{k+1}(u)=\left(\frac{\mathrm{}^2}{4}\frac{\text{d}^3}{\text{d}t_0^3}u\frac{\text{d}}{\text{d}t_0}\frac{1}{2}\frac{\text{d}u}{\text{d}t_0}\right)R_k(u),R_0=1,R_1=\frac{u}{2}.$$ . In the semiclassical limit (40) reduces to $$\underset{k=0}{\overset{m}{}}c_kt_ku^k=0\mathrm{or}_{\mathrm{}}\frac{\text{d}\mathrm{\Omega }}{2\pi }\frac{V^{}(\mathrm{\Omega },t)}{\sqrt{\mathrm{\Omega }+u(t)}}=0,$$ (41) where the $`c_k`$’s are the power series coefficients of $`(1+x)^{1/2}`$. In this semiclassical approximation $`u(t)=\sqrt{\mu (t)}`$ where $`\mu `$ is the effective cosmological constant of 2d Euclidean quantum gravity coupled to a $`(2,2m1)`$ conformal field theory, expressed via the string equation as a function of the coupling constants $`t_i`$. A choice of coupling constants $`t_i`$ is called a choice of closed string background and from (38) it is seen that the semiclassical disk amplitude can be written as $$W(x,t)=\frac{1}{2}\left(\underset{l=0}{\overset{m}{}}t_lP_l(x,u)\right)\sqrt{x+u},$$ (42) where the polynomial multiplying $`t_l`$ is $$P_l(x,u)=\underset{i=0}{\overset{l1}{}}c_ku^kx^{l1k}.$$ (43) Similar remarks applies to $`S(x,t)`$ which has a similar representation as (42) with polynomials $`\stackrel{~}{P}_l(x,u)`$ one degree higher than $`P_l(x,u)`$. Differentiating $`S(x,t)`$ w.r.t. $`t_0`$ reproduces precisely the FZZT amplitude for an insertion of the primary operator of the most negative (gravitationally dressed) scaling dimension <sup>2</sup><sup>2</sup>2Differentiating $`S`$ or $`W`$ after the other coupling constants $`t_i`$ produce expressions, also only depending on $`u`$ and $`q`$, with no explicit reference to $`t_i`$. However, in general the $`t_i`$’s cannot for $`0<im2`$ be identified directly with coupling constants related to the primary operators of scaling dimension $`(m2+i)/2`$ due to operator mixing with the (gravitational) descendents of the operator corresponding to $`t_0`$., $`\mathrm{\Delta }_{r=1,s=m1}=(m2)/2`$: $$\frac{\text{d}S(x,t)}{\text{d}t_0}=\sqrt{x+u}.$$ (44) We note that this formula implies that $`S(x,t)`$ $`=`$ $`{\displaystyle _{t_0(x)}^{t_0}}\text{d}t_0^{}\sqrt{x+u(t_0^{},t_1,\mathrm{})}`$ (45) $`W(x,t)`$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle _{t_0(x)}^{t_0}}\text{d}t_0^{}{\displaystyle \frac{1}{\sqrt{x+u(t_0^{},t_1,\mathrm{})}}},`$ (46) where $`u(t_0(x))=x`$. A full non-perturbative treatment involves the following steps: first solve the string equation (40) for $`u`$. Then solve the Schrödinger equation (13) for $`\psi (\tau ,x)`$ and finally construct the projector (18). The Wigner transform either of $`\psi (\tau ,x)`$ or the projector $`K(x,y;\tau )`$ then provides us with variables $`(q,p)`$ which can be viewed as phase space variables which lives on a non-commutative phase space defined by the projector as described above. Let us now show that in a naive semiclassical ($`\mathrm{}g_s0`$) limit we obtain formally the classical phase space $`p=W(q)`$, i.e. for the the choice of $`t_k`$’s corresponding to a conformal background precisely (1). Let us write the solution to (13) as: $$\psi (t_0,x)=\sqrt{R(x,u)}\text{e}^{\pm S(x,t_0)},\frac{S}{t_0}=\frac{1}{R}.$$ (47) The (13) can be written as $$\frac{\mathrm{}^2}{4}\left(\left(\frac{\text{d}R}{\text{d}t_0}\right)^22R\frac{\text{d}^2R}{\text{d}t_0^2}\right)+(x+u(t_0))R^2=1.$$ (48) Eqs. (47) and (48) is one version of the WKB-equations<sup>3</sup><sup>3</sup>3Note that if we differentiate (48) with respect to $`t_0`$ we obtain $$\left(\frac{\mathrm{}^2}{4}\frac{\text{d}^3}{\text{d}t_0^3}(x+u)\frac{\text{d}}{\text{d}t_0}\frac{1}{2}\frac{\text{d}u}{\text{d}t_0}\right)R=0,$$ which shows that $`R(x,u)`$ is the generating functional for the Gelfand-Dikii polynomials: $$R(x,u)=\underset{k=0}{\overset{\mathrm{}}{}}R_k(u)x^{k1/2}.$$ It also follows from the definition that $`R(x,u)`$ is the diagonal element of the resolvent: $`t_0|(Qx)^1|t_0`$., which are equivalent to the Schrödinger equation (13). In the WKB approximation where only the lowest order terms in $`\mathrm{}`$ is kept in $`R`$ and $`S`$ we have from (48) $$R(x,u)=\frac{1}{\sqrt{x+u}},$$ (49) and we obtain $`S`$ from (47) by integrating from the classical turning point, i.e. the point $`t_0(x)`$ where $`u(t_0(x),t_1,\mathrm{})=x`$, $$S(x,t)=_{t_0(x)}^{t_0}\text{d}t_0^{}\sqrt{x+u(t_0^{},t_1,\mathrm{})}.$$ (50) Comparing with (45) we see explicitly that the WKB approximation indeed reproduces the exponential of the semiclassical disk amplitude, i.e. the partition function of a single FZZT brane, as expected from the general discussion above. ### 3.1 The naive limit We can now use this WKB approximation to calculate the Wigner transform: $$f_\psi (q,p)=\frac{\text{d}x}{2\pi \mathrm{}}\text{e}^{ipx/\mathrm{}}\psi ^{}(qx/2)\psi (q+x/2).$$ (51) We emphasize that this Wigner transformed function is different from the ordinary Wigner function corresponding to $`\psi _x(t_0)`$ considered as an eigenfunction to the Schrödinger operator $`Q`$ with eigenvalue $`x`$. We perform the transformation in $`x`$, not in $`t_0`$, which is the variable appearing in $`Q`$. Since $`\mathrm{}0`$ it is tempting (but not entirely correct as will be discussed shortly) to change integration variable $`x=\mathrm{}y`$ and use the following approximations in the integrand: $$S(q+\mathrm{}y/2)=S(q)+\frac{\mathrm{}y}{2}\frac{S(q)}{q}+O(\mathrm{}^2),$$ (52) and $`R(q+\mathrm{}y/2,u)R^{}(q\mathrm{}y/2)=|R(q,u)|^2+O(\mathrm{}^2)`$. If we first consider $`q`$ which is in the ”classically allowed region” (where $`S(q)`$ is imaginary), and use eq. (52) in eq. (51) we obtain $$f_\psi (q,p)=|R(q,u(t))|\delta \left(p^2+\left(\frac{S(q,t)}{q}\right)^2\right)=\delta (t_0t_0^i(q,p)),$$ (53) where $`t_0^i(q,p)`$ is defined as the value of $`t_0`$ where the argument of the $`\delta `$-function is zero for given, fixed values of $`p`$ and $`q`$. In the “classical forbidden region” (where $`S(q)`$ is real) we obtain in the same approximation zero because the contribution is exponentially suppressed by $`\mathrm{exp}(2S(q)/\mathrm{})`$ and $`\mathrm{}0`$. From eq. (53) we finally obtain for the Wigner-transformed projector (see (26) and (18)) $$f_\rho (q,p)=\theta \left(p^2+\left(\frac{S(q,t)}{q}\right)^2\right).$$ (54) In the above semiclassical approximation we have thus seen that the Wigner transform gives us the classical phase-space (2), provided we make the identification $`Wip`$. This identification is indeed natural since the $`W`$ appearing in (2) should be viewed as a classical representative of the $`P`$ appearing in (14), i.e. a representation of the operator $`d/dx`$. Clearly, by construction, the $`p`$ appearing in the Wigner transform is a “real” $`p`$, related to $`id/dx`$. ### 3.2 The WKB Wigner function The approximation (52) does not commute unfortunately with the integration in (51). Thus the results (53) and (54) are not correct, although they indeed lead to the correct semiclassical expectation values of nice observables like $`\widehat{x}^n`$. A saddle-point calculation of (51) is however quite reliable except when the values of $`q`$ and $`p`$ are very close to the classical phase space orbit defined by (54). From (51) we obtain the saddle point equation for $`x`$: $$W(q+x_0/2)W^{}(qx_0/2)=2ip$$ (55) In the classically allowed region (where $`q+u(t)<0`$) $`W(q)`$ is imaginary and eq. (55) allows for ordinary quantum mechanical systems a standard solution (see ) since the phase space curve is concave (see fig. 1), as emphasized in . The saddle-point solution is $$f_\psi (q,p)=c^2(F(q,p)\mathrm{exp}\left(\frac{i}{\mathrm{}}A(q,p)\right)+\mathrm{c}.\mathrm{c}.)$$ (56) where $`c^2`$ is a normalization factor irrelevant for the present discussion. The prefactor $`F(q,p)`$ is $$F(q,p)=\frac{\sqrt{R(q+\frac{x_0}{2},t)R^{}(q\frac{x_0}{2},t)}}{\sqrt{W_{x_0}^{}(q+\frac{x_0}{2})+W_{x_0}^{}{}_{}{}^{}(q\frac{x_0}{2})}},$$ (57) where $`x_0`$ is the solution to (55) and $`W_{x_0}^{}`$ denotes the derivative after $`x_0`$. Finally $`A(q,p)`$ is the value of the exponent in (53) for the solution $`x_0`$ of (55), i.e. $$A(q,p)=px_0+S(q+\frac{x_0}{2})+S(q\frac{x_0}{2}).$$ (58) For a concave phase space orbit $`A`$ and $`(q,p)`$ in the ”interior”, $`A`$ clearly has the interpretation of the area shown in fig. 1. If $`W`$ is imaginary then $`x_0`$ is also a solution. Eq. (56) includes both saddle points via the complex conjugate term and $`f_\psi (q,p)`$ is real and oscillating. Note that these oscillations do not disappear in the limit $`\mathrm{}0`$. For an ordinary quantum mechanical system with $`(q,p)`$ outside the classically allowed region in phase space corresponding to a certain energy eq. (55) will have a complex solution $`x_0`$ and $`f_\psi (q,p)`$ will fall off exponentially. If eq. (55) holds for values of $`(q,p)`$ where $`x_00`$, i.e. when we approach the ”classical phase space” defined by (54), we see that (56) diverges due to the divergence of the prefactor $`F(q,p)`$ for $`x_00`$. This is not related to the question of the reliability of the WKB approximation of $`\psi (t_0,x)`$ at the classical turning point, but just to a failure of the quadratic saddle point calculation of the Wigner integral and one can improve the calculation by the method of uniform approximation, which involves in a natural way the Airy functions. This was done in where a closed expression was given, which is not only valid near the classical turning points but in the whole phase space, and which reduces to (56) away from the classical orbit: $$f_\psi ^{wkb}(q,p)=\frac{c^2}{\mathrm{}^{2/3}}F(q,p)A^{\frac{1}{6}}(q,p)\mathrm{Ai}\left(\left[\frac{3A(q,p)}{2\mathrm{}}\right]^{2/3}\right)$$ (59) where $`c`$ is a normalization constant independent of $`\mathrm{}`$, irrelevant for the present discussion. The Wigner WKB-function is a good approximation to the exact result for small $`\mathrm{}`$, in the sense that it incorporates already the improvements to the simple WKB approximation present in the so-called uniform semiclassical approximation even if the starting point was the simple WKB-approximation for $`\psi `$ . In particular, it is finite at the classical orbit except in exceptional cases discussed in . Again, the non-disappearance of $`f_\psi (q,p)`$ in the classically allowed region but away from the classical orbit is quite manifest in the expression (59). The following features are noteworthy: (1) the peak of $`f_\psi (q,p)`$ grows as $`\mathrm{}^{2/3}`$ as $`\mathrm{}0`$. Thus $`f_\psi (q,p)`$ a priori had a chance to reproduce the $`\delta `$-function at the classical orbit (53). However the Airy function oscillates in this region and these oscillations also diverge in the limit $`\mathrm{}0`$ although slightly weaker. Folded with sufficiently smooth functions these oscillations average out, and for ”observables” $`O(q,p)`$ corresponding to such functions the $`\mathrm{}0`$ limit of $`f_\psi (q,p)`$ will indeed be given by (53) (see for discussion and examples). (2) Since the peak of the Airy function is not at zero it follows from (59) that the ”fuzzy classical orbit” is actually displaced a distance proportional to $`\mathrm{}^{2/3}`$ and the spread of this fuzzy orbit is also of order $`\mathrm{}^{2/3}`$. ### 3.3 Beyond the semiclassical limit (almost) The simplest example is provided by the (2,1) minimal string model ($`m=1`$). In this case we have simply $`u(t)=t_0`$. Thus we can explicitly solve the system both in the semiclassical limit and completely non-perturbatively in $`\mathrm{}`$. Let us first state form of the WKB-approximation above. Let us denote $`t_0`$ by $`E`$, the ”classical energy”. Thus we have on a classical orbit $$p_{cl}^2+q_{cl}=E,\frac{3}{2}A(q,p)=(E(p^2+q))^{3/2},$$ (60) while $`F(q,p)A^{1/6}(q,p)`$, and we obtain from (59) $$f_\psi ^{wkb}(q,p)\frac{1}{\mathrm{}^{2/3}}\mathrm{Ai}\left(\frac{p^2+q+t_0}{\mathrm{}^{2/3}}\right).$$ (61) The exact solution was already cited earlier (31) and its Wigner transform is $$f_\psi (q,p)=\frac{1}{\pi \mathrm{}^{2/3}}\mathrm{Ai}\left(\frac{p^2+(q+t_0)}{\mathrm{}^{2/3}}\right)$$ (62) Maybe not surprisingly the Airy-function uniformized WKB approximation is exact for the linear potential! However, it suggests that many of the lessons one can draw from (59) might well be valid beyond the semiclassical approximation. At least here we have seen that in the context of the simplest non-critical string theory it includes a summation over all genera. Given the explicit expression (32) we can calculate the Wigner transform $`f_\rho (q,p)`$ of the projector $`K`$. The projector $`f_\rho (q,p)`$ is better behaved than $`f_\psi (q,p)`$ w.r.t. the classical limit. As shown in (34)–(35) the projector $`f_\rho (q,p)`$ indeed approaches the naive limit (53) for $`\mathrm{}0`$. This argument can clearly be extended to the general WKB Wigner transform given by (59). Integrating with respect to $`t_0`$ will produce additional factors of $`\mathrm{}`$ which will suppress the oscillations in the ”classically allowed region” in the limit $`\mathrm{}0`$. ## 4 Discussion Non-critical string theory has always been a useful laboratory for the study new ideas in string theory. Most recently it has received renewed attention after it was realized that the concepts of branes and open-closed string duality could be analyzed using matrix model technology. In this connection a target space interpretation of the the $`c<1`$ non-critical strings was suggested, target space being identified with a Riemann surface which was explicitly constructed in the conformal background for a $`(k,l)`$ non-critical string in the semiclassical approximation. One of the virtues of non-critical string theory is that it, at least to some extent, offers us a non-perturbative definition of string theory. In particular for the $`(2,2m1)`$ non-critical string theories, where $`m`$ is odd, this is relatively concrete: one can define a theory which should represent the non-perturbative sum over all genera in the sense that one can actually solve the string equation and find a “physically acceptable” string susceptibility function $`u(t)`$ well defined on the whole real $`t_0`$-axis . One problem arises in this connection: the elaborate structure of target space seemingly present semiclassically disappears completely and target space reduces to the complex plane. It seems a quite unsatisfactory situation if we extrapolate back to critical string theory. It would imply that we know nothing about target space unless we are able to perform the sum over all genera in critical string theory. It prompted us to look for another target space interpretation of the non-critical strings. We were here inspired by the fact that in $`c=1`$ non-critical string theory (where the notion of target space is more clear than for non-critical string theories with $`c<1`$) the concept of a non-commutative phase space appears naturally . In this note we have shown that noncommutative phase space indeed appears in a natural way in non-critical string theory. The Wigner function associated with the projector which defines the non-commutative phase space reduces in the semiclassical limit precisely to (54). It also makes it possible to avoid taking about Riemann surfaces altogether, if wanted. Target space becomes associated with a (real) coordinate $`q`$ to which we can associate a noncommutative ”momentum” $`p`$. Since we consider $`c<1`$ we have no time evolution in this target space but orbits of constant ”energy” makes sense and some aspects of the noncommutative nature of the phase space survive even semiclassically since the WKB approximation precisely defines the semiclassical FZZT brane. Let us perhaps contrast the appearance of noncommutative geometry in these theories with the earlier studies of noncommutativity in the context of backgrounds with nonzero $`B_{NSNS}`$ and the resulting open string theories . There the noncommutativity was a property of a particular background and could be turned off at will, while here it is intrinsically linked to the string coupling constant. Let us note that the same noncommutative geometrical structure appears naturally also in other contexts in string theory. $`c=1`$ strings also have a free fermion formulation with a filled Fermi sea whose classical boundary is the curve $`p^2q^2\mu =0`$. In an analogous fashion the quantum description will again be the noncommutative plane with a projection operator which represents a fuzzy noncommutative version of the Fermi sea. Since the $`c<1`$ and $`c=1`$ string theories are closely associated with B-model topological strings on Calabi-Yau’s of the form $`zwH(p,q)=0`$, the above results also suggest that similar noncommutative structures should also appear in that context. It would be also interesting to explore the issues of noncommutativity in the context of the $`\frac{1}{2}`$-BPS sector in the AdS/CFT correspondence which on the classical level is mapped to a free fermion description and the corresponding classical geometry is encoded by the geometry of the Fermi sea . Finally on a much more speculative level let us mention A-model topological strings formulated through the sum over 3D partitions . The blown-up Calabi-Yau’s (associated to toric diagrams $``$ 3D partitions) are represented by a lattice of points in an octant. The missing points are very naturally identified with harmonic oscillator eigenfunctions which are projected out<sup>4</sup><sup>4</sup>4E.g. such an interpretation for the toric diagram of $`^1`$ yields the fuzzy sphere.. The melting crystal partition function would then have an interpretation as a summation over (admissible) quantum geometries represented by projectors. It would be fascinating to understand more deeply the interrelations between all these pictures. #### Acknowledgment RJ was supported in part by KBN grants 2P03B08225 (2003-2006) and 1P03B02427 (2004-2007). Note added: As this paper was being finished the work appeared which also addresses the quantum phase space of minimal strings.
warning/0506/math0506532.html
ar5iv
text
# Interpolation in the noncommutative Schur-Agler class ## 1. Introduction The classical setting. By way of introduction we recall the classical Schur class $`𝒮`$ of analytic functions mapping the unit disk $`𝔻`$ into the closed unit disk $`\overline{𝔻}`$. The operator-valued Schur class $`𝒮(𝒰,𝒴)`$ consists, by definition, of analytic functions $`F`$ on $`𝔻`$ with values $`F(z)`$ equal to contraction operators between two Hilbert spaces $`𝒰`$ and $`𝒴`$. In what follows, the symbol $`(𝒰,𝒴)`$ stands for the algebra of bounded linear operators mapping $`𝒰`$ into $`𝒴`$, and we often abbreviate $`(𝒰,𝒰)`$ to $`(𝒰)`$. The class $`𝒮(𝒰,𝒴)`$ admits several remarkable characterizations. In particular any such function $`F(z)`$ can be realized in the form $$F(z)=D+zC(IzA)^1B$$ (1.1) where the connecting operator (or colligation) $$𝐔=\left[\begin{array}{cc}A& B\\ C& D\end{array}\right]:\left[\begin{array}{c}\\ 𝒰\end{array}\right]\left[\begin{array}{c}\\ 𝒴\end{array}\right]$$ is unitary, and where $``$ is some auxiliary Hilbert space (the internal space for the colligation). From the point of view of system theory, the function (1.1) is the transfer function of the linear system $$𝚺=𝚺(𝐔):\{\begin{array}{ccc}\hfill x(n+1)& =& Ax(n)+Bu(n)\hfill \\ \hfill y(n)& =& Cx(n)+Du(n)\hfill \end{array}.$$ It is also well known that the Schur-class functions satisfy a von Neumann inequality: if $`F𝒮(𝒰,𝒴)`$ and $`T(𝒦)`$ satisfies $`T<1`$, then $`F(T)`$ is a contraction operator ($`F(T)1`$), where $`F(T)`$ is defined by $$F(T)=\underset{n=0}{\overset{\mathrm{}}{}}F_nT^n(𝒰𝒦,𝒴𝒦)\text{if}F(z)=\underset{n=0}{\overset{\mathrm{}}{}}F_nz^n.$$ There is also a well-developed interpolation theory for the classical Schur class. One convenient formalism which encodes classical Nevanlinna-Pick and Carathéodory-Fejér interpolation (see e.g. ) proceeds as follows. Making use of power series expansions one can introduce the left and the right evaluation maps $$F^L(T_L)=\underset{n=0}{\overset{\mathrm{}}{}}T_L^nF_n\text{and}F^R(T_R)=\underset{n=0}{\overset{\mathrm{}}{}}F_nT_R^n,$$ (1.2) which make sense for $`F𝒮(𝒰,𝒴)`$ and for every choice of strictly contractive operators $`T_L(𝒴)`$ and $`T_R(𝒰)`$. One can then formulate an interpolation problem with the data sets consisting of two Hilbert spaces $`𝒦_L`$ and $`𝒦_R`$ and operators $`T_L(𝒦_L),T_R(𝒦_R),X_L(𝒴,𝒦_L),`$ $`Y_L(𝒰,𝒦_L),X_R(𝒦_R,𝒴),Y_R(𝒦_R,𝒰)`$ as follows: ###### Problem 1.1. Given the data as above, find necessary and sufficient conditions for existence of a function $`S𝒮(𝒰,𝒴)`$ such that $$\left(X_LS\right)^L(T_L)=Y_L\text{and}\left(SY_R\right)^R(T_R)=X_R.$$ (1.3) The answer is well known: Problem 1.3 has a solution if and only if there exists a positive semidefinite operator $`P(𝒦_L𝒦_R)`$ subject to the Stein identity $$M^{}PMN^{}PN=X^{}XY^{}Y$$ where $$M=\left[\begin{array}{cc}I_{𝒦_L}& 0\\ 0& T_R\end{array}\right],N=\left[\begin{array}{cc}T_L& 0\\ 0& I_{𝒦_R}\end{array}\right],X=\left[\begin{array}{cc}X_L^{}& X_R\end{array}\right],Y=\left[\begin{array}{cc}Y_L^{}& Y_R\end{array}\right].$$ (1.4) Multivariable extensions. Multivariable generalizations of these and many other related results have been obtained recently; one very general formulation introduced (see ) proceeds as follows. Let $`𝐐`$ be a $`m\times k`$ matrix-valued polynomial $$𝐐(z)=\left[\begin{array}{ccc}𝐪_{11}(z)& \mathrm{}& 𝐪_{1k}(z)\\ \mathrm{}& & \mathrm{}\\ 𝐪_{m1}(z)& \mathrm{}& 𝐪_{mk}(z)\end{array}\right]:^d^{m\times k}$$ (1.5) and let $`𝒟_𝐐^d`$ be the domain defined by $$𝒟_𝐐=\{z^d:𝐐(z)<1\}.$$ (1.6) For $`𝒰`$ and $`𝒴`$ two separable Hilbert spaces, in analogy with the classical case it is natural to define the Schur class for the domain $`𝒟_𝐐`$ as the class $`𝒮_𝐐(𝒰,𝒴)`$ of holomorphic $`(𝒰,𝒴)`$-valued functions $`S`$ on $`𝒟_𝐐`$ such that $`S(z)1`$ for all $`z𝒟_𝐐`$. We say that an $`S𝒮_𝐐(𝒰,𝒴)`$ satisfies the $`𝐐`$-von Neumann inequality over $`𝒟_𝐐`$ if $`S(T_1,\mathrm{},T_d)1`$ for all commuting tuples $`(T_1,\mathrm{},T_d)`$ of operators on a Hilbert space $`𝒦`$ with $`𝐐(T_1,\mathrm{},T_d)<1`$. (Here the fact that $`𝐐(T_1,\mathrm{},T_d)<1`$ implies that the Taylor joint spectrum of $`(T_1,\mathrm{},T_d)`$ is contained in $`𝒟_𝐐`$, so one can use a tensored version of the Taylor functional calculus to define $`S(T_1,\mathrm{},T_d)`$—see .) We define the Schur-Agler class over $`𝒟_𝐐`$, denoted by $`𝒮𝒜_𝐐(𝒰,𝒴)`$, to consist of all $`S𝒮_𝐐(𝒰,𝒴)`$ which in addition satisfy the $`𝐐`$-von Neumann inequality over $`𝒟_𝐐`$. As was first understood for the tridisk case $`(𝐐(z_1,z_2,z_3)=\left[\begin{array}{ccc}z_1& 0& 0\\ 0& z_2& 0\\ 0& 0& z_3\end{array}\right]`$), it can happen that the containment $`𝒮𝒜_𝐐(𝒰,𝒴)𝒮_𝐐(𝒰,𝒴)`$ is strict. It is this smaller class $`𝒮𝒜_𝐐(𝒰,𝒴)`$ which has a characterization analogous to (1.1) and thereby can be interpreted as the set of transfer functions of some type of conservative linear system, namely (see ): an $`(𝒰,𝒴)`$-valued function analytic on $`𝒟_𝐐`$ belongs to the class $`𝒮𝒜_𝐐`$ if and only if there exists an auxiliary Hilbert space $``$ and a unitary operator $$U=\left[\begin{array}{cc}A& B\\ C& D\end{array}\right]:\left[\begin{array}{c}^p\\ 𝒰\end{array}\right]\left[\begin{array}{c}^q\\ 𝒴\end{array}\right]$$ such that $$S(z)=D+C\left(I_^p(𝐐(z))A\right)^1(𝐐(z))B.$$ (1.7) Note that special choices of $$𝐐(z)=\mathrm{diag}(z_1,\mathrm{},z_d)\text{and}𝐐(z)=\left[\begin{array}{cccc}z_1& z_2& \mathrm{}& z_d\end{array}\right]$$ (1.8) lead to the unit polydisk $`𝒟_𝐐=𝔻^d`$ and the unit ball $`𝒟_𝐐=𝔹^d`$ of $`^d`$, respectively. The classes $`𝒮𝒜_𝐐(𝒰,𝒴)`$ for these two generic cases have been known for a while. The polydisk setting was first presented by J. Agler in and then extended to the operator valued case in ; see also . The Schur-Agler functions on the unit ball appeared in and later in in connection with complete Nevanlinna-Pick kernels and in in connection with the study of commutative unitary dilations of commutative row contractions; the Schur-Agler class for the unit ball case has the extra structure that it can be identified with the unit ball of the space of operator-valued multipliers over the Arveson space (the reproducing kernel Hilbert space over the unit ball $`𝔹^d^d`$ with reproducing kernel $`k_d(z,w)=\frac{1}{1z,w}`$)—we refer to for a thorough review of the operator-valued case. The case when $`𝒟_𝐐`$ is the Cartesian product of unit balls (of arbitrary finite dimensions) was considered in . Schur-Agler-class functions on $`𝔻^d`$ and $`𝔹^d`$ arise as the transfer functions of Givone-Roesser (see ) and Fornasini-Marchesini (see ) systems, respectively, which satisfy an additional energy-balance relation (see ). In the general case, formula (1.7) can be interpreted as representing $`S`$ as the transfer function of a more general type of multidimensional conservative linear system (see \[15, Section 4\] for more detail). An interpolation problem similar to Problem 1.3 has been studied in . Interpolation conditions for this problem are the same as in (1.3) but $`T_L`$ and $`T_R`$ are now commuting $`d`$-tuples satisfying conditions $$𝐐(T_L)<1\text{and}𝐐(T_R)<1$$ (1.9) and definitions of the left and the right evaluation maps are more involved and rely on the Martinelli kernel (see ) of the Taylor functional calculus . Similarly to the one variable case, the problem has a solution if and only if there is a positive semidefinite operator $`P((𝒦_L)^m(𝒦_R)^k)`$ subject to the Stein identity $$\underset{j=1}{\overset{m}{}}M_j^{}PM_j\underset{\mathrm{}=1}{\overset{k}{}}N_{\mathrm{}}^{}PN_{\mathrm{}}=X^{}XY^{}Y$$ (1.10) where $`X`$ and $`Y`$ are the same as in (1.4) and $`M_j`$ and $`N_{\mathrm{}}`$ are certain operators depending on $`T_L`$ and $`T_R`$ respectively (see \[15, Theorem 1.4\]). The noncommutative setting. System theoretical aspects of the above ideas has been extended recently to noncommutative multidimensional linear systems of a certain structure. These systems, called structured noncommutative multidimensional linear systems or SNMLSs in ) have evolution along a free semigroup rather than along an integer lattice as is usually taken in work in multidimensional linear system theory, and the transfer function is a formal power series in noncommuting indeterminants rather than an analytic function of several complex variables. Furthermore, the transfer function of a conservative SNMLS satisfies a certain von Neumann type inequality which leads to the definition of a noncommutative Schur-Agler class associated with certain noncommutative analogues of the domains $`𝒟_𝐐`$ (but where $`𝐐`$ is restricted to be linear). The main result \[17, Theorem 5.3\] states that every noncommutative Schur-Agler function admits a unitary realization similar to (1.7). The purpose of the present paper is to study related interpolation problems of Nevanlinna-Pick type in the noncommutative Schur-Agler class. The precise definitions and constructions involve a certain type of graph (an “admissible graph” as defined below). Let $`\mathrm{\Gamma }`$ be a graph consisting of a set of vertices $`V=V(\mathrm{\Gamma })`$ and edges $`E=E(\mathrm{\Gamma })`$. An edge $`e`$ connects its source vertex $`s`$, denoted by $`s=𝐬(e)V`$, to its range vertex $`r`$, denoted by $`r=𝐫(e)V`$. Following , we say that $`\mathrm{\Gamma }`$ is admissible if it is a finite ($`V`$ and $`E`$ are finite sets) bipartite graph such that each connected component is a complete bipartite graph. The latter means that: 1. the set of vertices $`V`$ has a disjoint partitioning $`V=S\dot{}R`$ into the set of source vertices $`S`$ and range vertices $`R`$, 2. $`S`$ and $`R`$ in turn have disjoint partitionings $`S=\dot{}_{k=1}^KS_k`$ and $`R=\dot{}_{k=1}^KR_k`$ into nonempty subsets $`S_1,\mathrm{},S_K`$ and $`R_1,\mathrm{},R_K`$ such that, for each $`s_kS_k`$ and $`r_kR_k`$ (with the same value of $`k`$) there is a unique edge $`e=e_{s_k,r_k}`$ connecting $`s_k`$ to $`r_k`$ ($`𝐬(e)=s_k`$, $`𝐫(e)=r_k`$), and 3. every edge of $`\mathrm{\Gamma }`$ is of this form. If $`v`$ is a vertex of $`\mathrm{\Gamma }`$ (so either $`vS`$ or $`vR`$) we denote by $`[v]`$ the path-connected component $`p`$ (i.e., the complete bipartite graph $`p=\mathrm{\Gamma }_k`$ with set of source vertices equal to $`S_k`$ and set of range vertices equal to $`R_k`$ for some $`k=1,\mathrm{},K`$) containing $`v`$. Thus, given two distinct vertices $`v_1,v_2SR`$, there is a path of $`\mathrm{\Gamma }`$ connecting $`v_1`$ to $`v_2`$ if and only if $`[v_1]=[v_2]`$ and this path has length 2 if both $`v_1`$ and $`v_2`$ are either in $`S`$ or in $`R`$ and has length 1 otherwise. In case $`sS`$ and $`rR`$ are such that $`[s]=[r]`$, we shall use the notation $`e_{s,r}`$ for the unique edge having $`s`$ as source vertex and $`r`$ as range vertex: $$e_{s,r}E\text{ determined by }𝐬(e_{s,r})=s,𝐫(e_{s,r})=r.$$ Note that $`e_{s,r}`$ is well defined only for $`sS`$ and $`rR`$ with $`[s]=[r]`$. For an admissible graph $`\mathrm{\Gamma }`$, let $`_E`$ be the free semigroup generated by the edge set $`E`$ of $`\mathrm{\Gamma }`$. An element of $`_E`$ is then a word $`w`$ of the form $`w=e_N\mathrm{}e_1`$ where each $`e_k`$ is an edge of $`\mathrm{\Gamma }`$ for $`k=1,\mathrm{},N`$. We denote the empty word (consisting of no letters) by $`\mathrm{}`$. The semigroup operation is concatenation: if $`w=e_N\mathrm{}e_1`$ and $`w^{}=e_N^{}^{}\mathrm{}e_1^{}`$, then $`ww^{}`$ is defined to be $$ww^{}=e_N\mathrm{}e_1e_N^{}^{}\mathrm{}e_1^{}.$$ Note that the empty word $`\mathrm{}`$ acts as the identity element for this semigroup. On occasion we shall have use of the notation $`we^1`$ for a word $`w_E`$ and an edge $`eE`$; by this notation we mean $$we^1=\{\begin{array}{cc}w^{}\hfill & \text{if }w=w^{}e,\hfill \\ \text{undefined}\hfill & \text{otherwise.}\hfill \end{array}$$ (1.11) with a similar convention for $`e^1w`$. By $`w^{}`$ we mean $`e_1\mathrm{}e_N`$, the transpose of $`w=e_N\mathrm{}e_1`$. For each $`eE`$, we define a matrix $`I_{\mathrm{\Gamma },e}=[I_{\mathrm{\Gamma },e;s,r}]_{sS,rR}`$ (with rows indexed by $`S`$ and columns indexed by $`R`$) with matrix entries given by $$I_{\mathrm{\Gamma },e;s,r}=\{\begin{array}{cc}1\hfill & \text{ if }(s,r)=(𝐬(e),𝐫(e)),\hfill \\ 0\hfill & \text{ otherwise. }\hfill \end{array}$$ (1.12) We then define the *structure matrix* $`Z_\mathrm{\Gamma }(z)`$ associated with each admissible graph $`\mathrm{\Gamma }`$ to be the linear form in the noncommuting indeterminants $`z=(z_e:eE)`$ given by $$Z_\mathrm{\Gamma }(z)=\underset{eE}{}I_{\mathrm{\Gamma },e}z_e.$$ (1.13) The latter function is the noncommutative analogue of $`𝐐(z)`$ in (1.5). However, if we let the variables $`(z_e:eE)`$ in (1.13) commute, we pick up only special examples of the polynomial matrix functions $`𝐐(z)`$, as will be clear from the Examples below. ###### Example 1.2. Structure matrix for the noncommutative ball. In this case, we take the admissible graph $`\mathrm{\Gamma }^{FM}`$ (where the label “FM” refers to Fornasini-Marchesini for system-theoretic reasons explained in ) to be a complete bipartite graph having only one source vertex. Thus we take $`S^{FM}=\{1\},`$ and $`R^{FM}=E^{FM}=\{1,\mathrm{},d\}`$ with $`𝐬^{FM}(i)=1`$, $`𝐫^{FM}(i)=i`$, i.e., $`n=1,m=d`$. Thus we have $$I_{\mathrm{\Gamma }^{FM},i}=\left[\begin{array}{ccccccc}0& \mathrm{}& 0& 1& 0& \mathrm{}& 0\end{array}\right],$$ where 1 is located in the $`i`$-th slot. Thus, the structure matrix for the noncommutative ball case is given by $$Z_{\mathrm{\Gamma }^{FM}}(z)=\underset{i=1}{\overset{d}{}}I_{\mathrm{\Gamma }^{FM},i}z_i=\left[\begin{array}{ccc}z_1& \mathrm{}& z_d\end{array}\right].$$ Note that when the variables $`z_1,\mathrm{},z_d`$ commute, then the associated domain $`\{z=(z_1,\mathrm{},z_d):Z_{\mathrm{\Gamma }^{FM}}(z)<1\}`$ is the unit ball in $`^d`$. ###### Example 1.3. Structure matrix for the noncommutative polydisk. In this case, we take the admissible graph $`\mathrm{\Gamma }^{GR}`$ (where the label “GR” refers to Givone-Roesser for system-theoretic reasons explained in ) to have $`d`$ path-connected components with each path-connected component containing only one source and one range vertex. Thus, we take $`S^{GR}=R^{GR}=E^{GR}=\{1,\mathrm{},d\}`$ with $`𝐬^{GR}(i)=i`$, $`𝐫^{GR}(i)=i`$ and thus $`n=d=m`$. Then $`I_{\mathrm{\Gamma }^{GR},i}`$ is the $`d\times d`$ matrix with $`1`$ located at the $`(i,i)`$-th entry and with all other entries are zeros. Therefore, the structure matrix for the noncommutative Givone-Roesser case has the diagonal form $$Z_{\mathrm{\Gamma }^{GR}}(z)=\underset{i=1}{\overset{d}{}}z_iI_{\mathrm{\Gamma }^{GR},i}=\left[\begin{array}{ccc}z_1& & \\ & \mathrm{}& \\ & & z_d\end{array}\right].$$ If the variables $`z_1,\mathrm{},z_d`$ commute, then the associated domain $`\{z=(z_1,\mathrm{},z_d):Z_{\mathrm{\Gamma }^{GR}}(z)<1\}`$ is the unit polydisk in $`^d`$. ###### Example 1.4. Full matrix block structure matrix. In this case, we take $`\mathrm{\Gamma }^{\text{full}}`$ to be a general finite, complete bipartite graph. Thus we take $`S=\{1,\mathrm{},n\}`$, $`R=\{1,\mathrm{},m\}`$, and $`E=\{(i,j):iS,jR\}`$ with $`𝐬^{\text{full}}(i,j)=i`$, $`𝐫^{\text{full}}(i,j)=j`$ where $`d=nm`$. Then $`I_{\mathrm{\Gamma }^{\text{full}},(i,j)}`$ is the $`d\times d`$ matrix with $`1`$ located at the $`(i,j)`$-th entry and all other entries are zeros. Thus the structure matrix for this case has the full-block structure $$Z_{\mathrm{\Gamma }^{\text{full}}}(z)=\left[\begin{array}{ccc}z_{1,1}& \mathrm{}& z_{1,m}\\ \mathrm{}& & \mathrm{}\\ z_{n,1}& \mathrm{}& z_{n,m}\end{array}\right].$$ ###### Example 1.5. The general structure matrix. Suppose that the admissible graph $`\mathrm{\Gamma }`$ has path connected components $`\mathrm{\Gamma }_k`$ with source vertices $`S_k=\{(k,1),\mathrm{},(k,n_k)\}`$, range vertices $`R_k=\{(k,1),\mathrm{},(k,m_k)\}`$ and edge sets $`E_k=\{(k,i,j):1in_k,1jm_k\}`$ for $`k=1,\mathrm{},K`$. Define a graph $`\mathrm{\Gamma }`$ to have source vertex set $$S=_{k=1}^KS_k=\{(k,i):1kK,1in_k\},$$ range vertex set $$R=_{k=1}^KR_k=\{(k,j):1kK,1jm_k\}$$ and edge set $$E=_{k=1}^KE_k=\{(k,i,j):1kK,1in_k,1jm_k\}$$ with $`𝐬(k,i,j)=(k,i)`$, $`𝐫(k,i,j)=(k,j)`$ for $`(k,i,j)E`$. Then the associated structure matrix $`Z_\mathrm{\Gamma }(z)`$ is given by $$Z_\mathrm{\Gamma }(z)=\left[\begin{array}{ccc}Z_{\text{full},1}(z^1)& & \\ & \mathrm{}& \\ & & Z_{\text{full},K}(z^K)\end{array}\right]$$ where we let $`z^k`$ denote the $`(n_km_k)`$-tuple of variables $`z^k=(z_{k,i,j}:1in_k;1jm_k)`$ and where $$Z_{\text{full},k}(z^k)=\left[\begin{array}{ccc}z_{k,1,1}& \mathrm{}& z_{k,1,m_k}\\ \mathrm{}& & \mathrm{}\\ z_{k,n_k,1}& \mathrm{}& z_{k,n_k,m_k}\end{array}\right]$$ is as in Example 1.4 for $`k=1,\mathrm{},K`$. By the definition of an admissible graph as a graph with path-connected components equal to complete bipartite graphs, we see that Example 1.5 amounts to the general case. Thus, the case considered in the present framework corresponds (in the commutative setting) not to arbitrary polynomials (1.5), but just to homogeneous linear functions in which case, the corresponding domain $`𝒟_𝐐`$ is the Cartesian product of finitely many Cartan domains of type $`I`$. The proofs of realization and interpolation results in this particular case are not substantially easier; however, most of needed constructions can be expressed in terms of uniformly converging power series rather than the Vasilescu’s operator analogue of the Martinelli-Bochner kernel. Thus, the transfer to the noncommutative setting via noncommutative formal power series in this situation is much more clear. In what follows, $`(𝒰,𝒴)z`$ will stand for the space of formal power series $`F`$ of the form $$F(z)=\underset{v_E}{}F_vz^v,F_v(𝒰,𝒴)$$ (1.14) in noncommutative variables $`z=\{z_e:eE\}`$ indexed by the edge set $`E`$ of the admissible graph $`\mathrm{\Gamma }`$, with coefficients $`F_v`$ equal to bounded operators acting between Hilbert spaces $`𝒰`$ and $`𝒴`$. Here $`z^{\mathrm{}}=1`$ and $`z^w=z_{e_N}z_{e_{N1}}\mathrm{}z_{e_1}`$ if $`w=e_Ne_{N1}\mathrm{}e_1`$. Thus $$z^wz^w^{}=z^{ww^{}},z^wz_e=z^{we}\text{ for }w,w^{}_E\text{ and }eE.$$ On occasion we shall have need of multiplication on the right or left by $`z_e^1`$; we use the convention $$z^wz_e^1=\{\begin{array}{cc}z^{we^1}\hfill & \text{if }we^1_E\text{ is defined;}\hfill \\ 0\hfill & \text{if }we^1\text{ is undefined,}\hfill \end{array}$$ (1.15) where we use the convention (1.11) for the meaning of $`we^1`$. We use the obvious analogous convention to define $`z_e^1z^w`$. Let $`T=(T_e:eE)`$ be a collection of bounded, linear operators (not necessarily commuting) on some separable infinite-dimensional Hilbert space $`𝒦`$ (also indexed by the edge set $`E`$ of $`\mathrm{\Gamma }`$). We define an operator $`F(T):𝒰𝒦𝒴𝒦`$ by $`F(T):=\underset{N\mathrm{}}{lim}{\displaystyle \underset{v_E:|v|N}{}}F_vT^v`$ $`\text{ where }T^{\mathrm{}}=I_𝒦\text{ and }T^v=T_{e_N}\mathrm{}T_{e_1}\text{ if }v=e_N\mathrm{}e_1`$ (1.16) whenever the limit exists in the weak-operator topology. <sup>1</sup><sup>1</sup>1In the limit is taken in the norm-operator topology; the weak-operator topology is more convenient for our purposes here. In general there is no reason for the limit in (1.16) to exist; on the other hand if $`F`$ is a polynomial in $`z`$, its action on noncommutative tuples is well defined. Alternatively, if $`T`$ is a nilpotent tuple (so that $`F^v=0`$ once the length $`|v|`$ of $`v`$ is large enough), then the expression (1.16) is well defined. More generally, it is well defined if $`F(z)`$ is a rational formal power series and the tuple $`T`$ is in the domain of $`F(z)`$—see . Take the function $`Z_\mathrm{\Gamma }`$ as in (1.13), define (according to (1.16)) the operator $$Z_\mathrm{\Gamma }(T):=\underset{eE}{}I_{\mathrm{\Gamma };e}T_e(_{rR}𝒦,_{sS}𝒦)$$ and introduce the noncommutative structured ball $$_\mathrm{\Gamma }(𝒦)=\{T=(T_e)_{eE}:T_e(𝒦)\text{ for }eE\text{ and }Z_\mathrm{\Gamma }(T)<1\}.$$ (1.17) Now we are in position to define the noncommutative Schur-Agler class. ###### Definition 1.6. Given an admissible graph $`\mathrm{\Gamma }`$, a formal power series (1.14) is said to belong to the noncommutative Schur-Agler class $`𝒮𝒜_\mathrm{\Gamma }(𝒰,𝒴)`$ if, for each Hilbert space $`𝒦`$ and each $`T=(T_e)_{eE}_\mathrm{\Gamma }(𝒦)`$, the limit $$F(T)=\underset{N\mathrm{}}{lim}\underset{v_E:|v|N}{}F_vT^v$$ (1.18) exists in the weak-operator topology and defines a contractive operator $$F(T):𝒰𝒦𝒴𝒦,F(T)1.$$ We remark that, for the noncommutative polydisk setting of Example 1.3, Alpay and Kalyuzhnyĭ-Verbovetzkiĭ in show that it suffices to check that the expression (1.18) is a contraction only for $`T_\mathrm{\Gamma }(𝒦)`$ with $`𝒦`$ a Hilbert space of arbitrarily large but finite dimension. The noncommutative analogue of the unitary realization (1.7) for the Schur-Agler class $`𝒮𝒜_\mathrm{\Gamma }(𝒰,𝒴)`$ was obtained in . To formulate the result we shall need some additional notation and terminology. First, given a collection $`=\{_p:pP\}`$ of Hilbert spaces indexed by the set $`P`$ of path-connected components of $`\mathrm{\Gamma }`$, let $$Z_{\mathrm{\Gamma },}(z)=\underset{eE}{}I_{\mathrm{\Gamma },;e}z_e$$ (1.19) where $`I_{\mathrm{\Gamma },;e}:_{rR}_{[r]}_{sS}_{[s]}`$ is given via matrix entries $$[I_{\mathrm{\Gamma },;e}]_{s,r}=\{\begin{array}{cc}I_{_{[𝐬(e)]}}=I_{_{[𝐫(e)]}}& \text{ if }s=𝐬(e)\text{ and }r=𝐫(e),\\ 0& \text{otherwise.}\end{array}$$ Furthermore, let $`z^{}=(z_e^{}:eE)`$ be another system of noncommuting indeterminants; while $`z_ez_e^{}z_e^{}z_e`$ and $`z_e^{}z_e^{}^{}z_e^{}^{}z_e^{}`$ unless $`e=e^{}`$, we will use the convention that $`z_ez_e^{}^{}=z_e^{}^{}z_e`$ for all $`e,e^{}E`$. We also shall need the convention (1.15) to give meaning to expressions of the form $$z_e^1z^vz^v^{}z_e^1=(z^vz_e^1)(z_e^1z^v^{})=z^{ve^1}z^{e^1v^{}}.$$ For $`F(z)`$ of the form (1.14), we will use the convention that $$F(z)^{}=\left(\underset{v_E}{}F_vz^v\right)^{}:=\underset{v_E}{}F_v^{}z^v^{}=\underset{v_E}{}F_v^{}^{}z^v.$$ We also use the notation $$\mathrm{Row}_{xX}M_x=\left[\begin{array}{ccc}M_{x_1}& \mathrm{}& M_{x_N}\end{array}\right],\mathrm{Col}_{xX}M_x=\left[\begin{array}{c}M_{x_1}\\ \mathrm{}\\ M_{x_N}\end{array}\right]\text{ if }X=\{x_1,\mathrm{},x_N\}$$ for block row and column matrices with rows or columns indexed by the set $`X`$. ###### Theorem 1.7. Let $`F(z)`$ be a formal power series in noncommuting indeterminants $`z=(z_e:eE)`$ indexed by the $`E`$ of edges of the admissible graph $`\mathrm{\Gamma }`$ with coefficients $`F_v(𝒰,𝒴)`$ for two Hilbert spaces $`𝒰`$ and $`𝒴`$. The following are equivalent: 1. $`F`$ belongs to the noncommutative Schur-Agler class $`𝒮𝒜_\mathrm{\Gamma }(𝒰,𝒴)`$. 2. There exist a collection $`=\{_p:pP\}`$ of Hilbert spaces indexed by the set $`P`$ of path-connected components $`\mathrm{\Gamma }`$ and a unitary operator $$U=\left[\begin{array}{cc}A& B\\ C& D\end{array}\right]:\left[\begin{array}{c}_{sS}_{[s]}\\ 𝒰\end{array}\right]\left[\begin{array}{c}_{rR}_{[r]}\\ 𝒴\end{array}\right]$$ such that $$F(z)=D+C(IZ_{\mathrm{\Gamma },}(z)A)^1Z_{\mathrm{\Gamma },}(z)B$$ (1.20) where $`Z_{\mathrm{\Gamma },}`$ is defined in (1.19). 3. There exist a collection of Hilbert spaces $`=\{_p:pP\}`$ and a formal power series $$H(z)=\mathrm{Row}_{sS}H_s(z)(_{sS}_{[s]},𝒴)z$$ (1.21) so that $$I_𝒴F(z)F(z^{})^{}=H(z)\left(IZ_{\mathrm{\Gamma },}(z)Z_{\mathrm{\Gamma },}(z^{})^{}\right)H(z^{})^{}.$$ (1.22) 4. There exist a collection of Hilbert spaces $`=\{_p:pP\}`$ and a formal power series $$G(z)=\mathrm{Col}_{rR}G_r(z)(𝒰,_{rR}_{[r]})z$$ (1.23) so that $$I_𝒰F(z)^{}F(z^{})=G(z)^{}\left(IZ_{\mathrm{\Gamma },}(z)^{}Z_{\mathrm{\Gamma },}(z^{})\right)G(z^{}).$$ (1.24) 5. There exist a collection of Hilbert spaces $`=\{_p:pP\}`$ and formal power series $`H(z)`$ and $`G(z)`$ as in (1.21), (1.23) so that relations (1.22), (1.24) hold along with $$F(z)F(z^{})=H(z)\left(Z_{\mathrm{\Gamma },}(z)Z_{\mathrm{\Gamma },}(z^{})\right)G(z^{}).$$ (1.25) A representation of the form (1.20) with $`U=\left[\begin{array}{cc}A& B\\ C& D\end{array}\right]`$ is called a unitary realization for $`F`$, or, in more detail in the terminology from , a realization of $`F`$ as the transfer function for the conservative Structured Noncommutative Multidimensional Linear System $`𝚺=\{\mathrm{\Gamma },,𝒰,𝒴,𝐔\}`$ (see Section 2 for further details). Note that if $`F`$ is of the form (1.20), then representations (1.22), (1.24) and (1.25) are valid with $$H(z)=C\left(IZ_{\mathrm{\Gamma },}(z)A\right)^1\text{and}G(z)=\left(IAZ_{\mathrm{\Gamma },}(z)\right)^1B.$$ (1.26) Now we turn to the subject of the paper. We shall consider bitangential interpolation problems with the data set consisting of two Hilbert spaces $`𝒦_L`$ and $`𝒦_R`$, two tuples $`T_L=\{T_{L,e}:eE\}`$ and $`T_R=\{T_{R,e}:eE\}`$ of operators acting on $`𝒦_L`$ and $`𝒦_R`$ respectively, and bounded operators $$X_L:𝒴𝒦_L,Y_L:𝒰𝒦_L,X_R:𝒦_R𝒴,Y_R:𝒦_R𝒰.$$ The pair $`(T_L,X_L)`$ will be said to be left admissible (with respect to the Schur-Agler class $`𝒮𝒜_\mathrm{\Gamma }(𝒰,𝒴)`$) if the left-tangential evaluation map (with operator argument) $`H(X_LH)^L(T_L)`$ given by $$(X_LH)^L(T_L)=\underset{v_E}{}T_L^v^{}X_LH_v$$ (1.27) is well-defined (with convergence of the infinite series in the weak-operator topology) whenever $`H(z)=_{v_E}H_vz^v`$ is a formal power series of the form (1.21) appearing in the representation (1.22) for a Schur-Agler class formal power series $`F(z)=_{v_E}F_vz^v𝒮𝒜_\mathrm{\Gamma }(𝒰,𝒴)`$. Whenever this is the case, from the identity $`F(z)=D+H(z)Z_{\mathrm{\Gamma },}(z)B`$ we read off that then the left-tangential map is also well-defined on the associated Schur-Agler class formal power series $`F(z)`$: $$(X_LF)^L(T_L)=\underset{v_E}{}T_L^v^{}X_LF_v=X_LD+\underset{eE}{}T_{L,e}[(X_LH_{𝐬(e)})^L(T_L)]B_{𝐫(e)}.$$ (1.28) Similarly, we say that the pair $`(Y_R,T_R)`$ is right admissible (with respect to the Schur-Agler class $`𝒮𝒜_\mathrm{\Gamma }(𝒰,𝒴)`$) if the right-tangential evaluation map (with operator argument) $`G(GY_R)^R(T_R)`$ given by $$(GY_R)^R(T_R)=\underset{v_E}{}G_vY_RT_R^v^{}$$ (1.29) exists (with convergence of the infinite series in the weak-operator topology) whenever $`G(z)=_{v_E}G_vz^v`$ is a formal power series of the form (1.23) appearing in the representation (1.24) for a Schur-Agler class formal power series $`F(z)=_{v_E}F_vz^v𝒮𝒜_𝒢(𝒰,𝒴)`$. Using the identity $`F(z)=D+CZ_{\mathrm{\Gamma },}(z)G(z)`$ we then see that the right-tangential evaluation map (with operator argument) is well-defined on the associated Schur-Agler class formal power series $`F(z)`$ as well: $$(FY_R)^R(T_R)=\underset{v_E}{}F_vY_RT_R^v^{}=DY_R+\underset{eE}{}C_{𝐬(e)}[(G_{𝐫(e)}Y_R)^R(T_R)]T_{R,e}.$$ (1.30) The connections between left and right point evaluation with operator argument given by (1.27) and (1.29) versus the tensor-product functional calculus given by (1.18) will be discussed in Section 3. We say that the data set $$𝒟=\{T_L,T_R,X_L,Y_L,X_R,Y_R\},$$ (1.31) is admissible (with respect to $`𝒮𝒜_\mathrm{\Gamma }(𝒰,𝒴)`$) if $`(T_L,X_L)`$ is left admissible and $`(Y_R,T_R)`$ is right admissible. We shall give examples and further details on admissible interpolation data sets in Section 3 below. Given an admissible interpolation data set (1.31), the formal statement of the associated bitangential interpolation problem is: ###### Problem 1.8. Find necessary and sufficient conditions for existence of a power series $`F𝒮𝒜_\mathrm{\Gamma }(𝒰,𝒴)`$ such that $$\left(X_LF\right)^L(T_L)=Y_L\text{and}\left(FY_R\right)^R(T_R)=X_R.$$ (1.32) To formulate the solution criterion we need some additional notation. Let $`𝜹_{s,s^{}}`$ be the Kronecker delta function $$𝜹_{s,s^{}}=\{\begin{array}{cc}1\hfill & \text{if }s=s^{},\hfill \\ 0\hfill & \text{otherwise}.\hfill \end{array}$$ For $`sS`$ and $`rR`$, define operators $`E_{L,s}`$ $`=`$ $`\mathrm{Col}_{s^{}S:[s^{}]=[s]}𝜹_{s,s^{}}I_{𝒦_L}:𝒦_L{\displaystyle \underset{s^{}S:[s^{}]=[s]}{}}𝒦_L,`$ (1.33) $`E_{R,r}`$ $`=`$ $`\mathrm{Col}_{r^{}R:[r^{}]=[r]}𝜹_{r,r^{}}I_{𝒦_R}:𝒦_R{\displaystyle \underset{r^{}R:[r^{}]=[r]}{}}𝒦_R,`$ (1.34) $`\stackrel{~}{N}_r(T_L)`$ $`=`$ $`\mathrm{Col}_{s^{}S:[s^{}]=[r]}T_{L,e_{s^{},r}}^{}:𝒦_L{\displaystyle \underset{s^{}S:[s^{}]=[r]}{}}𝒦_L,`$ (1.35) $`\stackrel{~}{M}_s(T_R)`$ $`=`$ $`\mathrm{Col}_{r^{}R:[r^{}]=[s]}T_{R,e_{s,r^{}}}:𝒦_R{\displaystyle \underset{r^{}R:[r^{}]=[s]}{}}𝒦_R.`$ (1.36) Define also the operators $`M_s`$ $`=`$ $`M_s(T_R)=\left[\begin{array}{cc}E_{L,s}& 0\\ 0& \stackrel{~}{M}_s(T_R)\end{array}\right](sS),`$ (1.37) $`N_r`$ $`=`$ $`N_r(T_L)=\left[\begin{array}{cc}\stackrel{~}{N}_r(T_L)& 0\\ 0& E_{R,r}\end{array}\right](rR).`$ (1.38) ###### Theorem 1.9. There is a power series $`F𝒮𝒜_\mathrm{\Gamma }(𝒰,𝒴)`$ satisfying interpolation conditions (1.32) if and only if there exists a collection $`𝕂=\{𝕂_p:pP\}`$ of positive semidefinite operators $$𝕂_p((_{sS:[s]=p}𝒦_L)(_{rR:[r]=p}𝒦_R))$$ indexed by the set $`P`$ of path-connected components of $`\mathrm{\Gamma }`$, which satisfies the Stein identity $$\underset{sS}{}M_s^{}𝕂_{[s]}M_s\underset{rR}{}N_r^{}𝕂_{[r]}N_r=X^{}XY^{}Y,$$ (1.39) where $`M_s`$ and $`N_r`$ are the operators defined via formulas (1.37), (1.38) and where $$X=\left[\begin{array}{cc}X_L^{}& X_R\end{array}\right]\text{and}Y=\left[\begin{array}{cc}Y_L^{}& Y_R\end{array}\right].$$ (1.40) Let $`𝕂=\{𝕂_p:pP\}`$ be any collection of operators satisfying the conditions in Theorem 1.40. Let us represent these operators more explicitly as $$𝕂_p=\left[\begin{array}{cc}𝕂_{p,L}& 𝕂_{p,LR}\\ 𝕂_{p,LR}^{}& 𝕂_{p,R}\end{array}\right]$$ (1.41) where $$𝕂_{p,L}=\left[\mathrm{\Psi }_{s,s^{}}\right],𝕂_{p,R}=\left[\mathrm{\Phi }_{r,r^{}}\right],𝕂_{p,LR}=\left[\mathrm{\Lambda }_{s,r}\right]$$ (1.42) for $`s,s^{}S`$ and $`r,r^{}R`$ such that $`[s]=[s^{}]=[r]=[r^{}]=p`$ and with $$\mathrm{\Psi }_{s,s^{}}(𝒦_L),\mathrm{\Phi }_{r,r^{}}(𝒦_R),\mathrm{\Lambda }_{s,r}(𝒦_R,𝒦_L).$$ (1.43) It turns out that for every collection $`𝕂=\{𝕂_p:pP\}`$ of positive semidefinite operators satisfying (1.39), there is a solution $`F`$ of the bitangential interpolation Problem 1.8 such that, for some choice of associated functions $`H(z)`$ and $`G(z)`$ of the form (1.21) and (1.23) in representations (1.22), (1.24), (1.25), it holds that $`(X_LH_s)^L(T_L)\left[(X_LH_s^{})^L(T_L)\right]^{}`$ $`=\mathrm{\Psi }_{s,s^{}}\text{ for }s,s^{}S:[s]=[s^{}],`$ (1.44) $`(X_LH_s)^L(T_L)\left(G_rY_R\right)^R(T_R)`$ $`=\mathrm{\Lambda }_{s,r}\text{ for }sS;rR:[s]=[r],`$ (1.45) $`\left[(G_rY_R)^R(T_R)\right]^{}(G_r^{}Y_R)^R(T_R)`$ $`=\mathrm{\Phi }_{r,r^{}}\text{ for }r,r^{}R:[r]=[r^{}].`$ (1.46) Furthermore, it turns out that conversely, for every solution $`F`$ of Problem 1.8 with representations (1.22), (1.24), (1.25) (existence of these representations is guaranteed by Theorem 1.7), the operators $`𝕂_p`$ defined via (1.41)–(1.43) and (1.44)–(1.46) satisfy conditions of Theorem 1.40. These observations suggest the following modification of Problem 1.8 with the data set $$𝒟=\{T_L,T_R,X_L,Y_L,X_R,Y_R,\mathrm{\Psi }_{s,s^{}},\mathrm{\Phi }_{r,r^{}},\mathrm{\Lambda }_{s,r}\}.$$ (1.47) ###### Problem 1.10. Given the data $`𝒟`$ as in (1.47), find all power series $`F𝒮𝒜_\mathrm{\Gamma }(𝒰,𝒴)`$ satisfying interpolation conditions (1.32) and such that for some choice of associated functions $`H_s`$ and $`G_r`$ in the representations (1.22), (1.24), (1.25), the equalities (1.44)–(1.46) hold. In contrast to Problem 1.8, the solvability criterion for Problem 1.10 can be given explicitly in terms of the interpolation data. ###### Theorem 1.11. Problem 1.10 has a solution if and only if the operators $`𝕂_p`$ ($`pP`$) given by (1.41), (1.42) are positive semidefinite and satisfy the Stein identity (1.39). Moreover, there exist Hilbert spaces $`\stackrel{~}{\mathrm{\Delta }}`$ and $`\stackrel{~}{\mathrm{\Delta }}_{}`$, a collection of Hilbert spaces $`\widehat{}=\{\widehat{}_p:pP\}`$ indexed by set of path-connected components $`P`$ of $`\mathrm{\Gamma }`$, and a formal power series $$\mathrm{\Sigma }(z)=\left[\begin{array}{cc}\mathrm{\Sigma }_{11}(z)& \mathrm{\Sigma }_{12}(z)\\ \mathrm{\Sigma }_{21}(z)& \mathrm{\Sigma }_{22}(z)\end{array}\right]:\left[\begin{array}{c}𝒰\\ \stackrel{~}{\mathrm{\Delta }}_{}\end{array}\right]\left[\begin{array}{c}𝒴\\ \stackrel{~}{\mathrm{\Delta }}\end{array}\right]$$ from the noncommutative Schur-Agler class $`𝒮𝒜_\mathrm{\Gamma }(𝒰\stackrel{~}{\mathrm{\Delta }}_{},𝒴\stackrel{~}{\mathrm{\Delta }})`$ of the form $$\mathrm{\Sigma }(z)=\left[\begin{array}{cc}U_{22}& U_{23}\\ U_{32}& 0\end{array}\right]+\left[\begin{array}{c}U_{21}\\ U_{31}\end{array}\right](I_{\stackrel{~}{\mathrm{\Delta }}_{}}Z_{\mathrm{\Gamma },\widehat{}}(z)U_{11})^1Z_{\mathrm{\Gamma },\widehat{}}(z)\left[\begin{array}{cc}U_{12}& U_{13}\end{array}\right]$$ (1.48) with $$𝐔_0=\left[\begin{array}{ccc}U_{11}& U_{12}& U_{13}\\ U_{21}& U_{22}& U_{23}\\ U_{31}& U_{32}& 0\end{array}\right]:\left[\begin{array}{c}_{sS}\widehat{}_{[s]}\\ 𝒰\\ \stackrel{~}{\mathrm{\Delta }}_{}\end{array}\right]\left[\begin{array}{c}\widehat{}_{rR}\widehat{}_{[r]}\\ 𝒴\\ \stackrel{~}{\mathrm{\Delta }}\end{array}\right]$$ unitary and completely determined by the interpolation data set $`𝒟`$ so that $`F`$ is a solution of Problem 1.10 if and only if $`F`$ has the form $$F(z)=\mathrm{\Sigma }_{11}(z)+\mathrm{\Sigma }_{12}(z)\left(I_{\stackrel{~}{\mathrm{\Delta }}_{}}𝒯(z)\mathrm{\Sigma }_{22}(z)\right)^1𝒯(z)\mathrm{\Sigma }_{21}(z)$$ (1.49) for a power series $`𝒯(z)𝒮𝒜_\mathrm{\Gamma }(\stackrel{~}{\mathrm{\Delta }},\stackrel{~}{\mathrm{\Delta }}_{})`$. As a corollary we have the following less satisfactory parametrization of the set of all solutions of Problem 1.8. ###### Corollary 1.12. Suppose that we are given a noncommutative interpolation data set $`𝒟`$ as in (1.31) and let $`𝐊`$ be the set of all collections $`𝕂=\{𝕂_p:pP\}`$ of positive semi-definite operators $`𝕂_p((_{sS:[s]=p}𝒦_L)(_{rR:[r]=p}𝒦_R))`$ which satisfy the Stein identity (1.39). For each $`𝕂𝐊`$, let $$\mathrm{\Sigma }^𝕂(z)=\left[\begin{array}{cc}\mathrm{\Sigma }_{11}^𝕂(z)& \mathrm{\Sigma }_{12}^𝕂(z)\\ \mathrm{\Sigma }_{21}^𝕂(z)& \mathrm{\Sigma }_{22}^𝕂(z)\end{array}\right]:\left[\begin{array}{c}𝒰\\ \stackrel{~}{\mathrm{\Delta }}_{}^𝕂\end{array}\right]\left[\begin{array}{c}𝒴\\ \stackrel{~}{\mathrm{\Delta }}^𝕂\end{array}\right]$$ be the characteristic function associated with $`𝕂`$ as in Theorem 1.11. Then the formal power series $`F(z)=_{v_E}T_vz^v`$ with coefficients $`T_v(𝒰,𝒴)`$ is a solution of Problem 1.8 if and only if there is a choice of $`𝕂𝐊`$ and a free-parameter formal power series $`𝒯(z)`$ in the Schur-Agler class $`𝒮𝒜_\mathrm{\Gamma }(\stackrel{~}{\mathrm{\Delta }}^𝕂,\stackrel{~}{\mathrm{\Delta }}_{}^𝕂)`$ so that $`F(z)`$ has the form $$F(z)=\mathrm{\Sigma }_{11}^𝕂(z)+\mathrm{\Sigma }_{12}^𝕂(z)(I_{\stackrel{~}{\mathrm{\Delta }}_{}^𝕂}𝒯(z)\mathrm{\Sigma }_{22}^𝕂(z))^1𝒯(z)\mathrm{\Sigma }_{21}^𝕂(z).$$ There has been some work on noncommutative interpolation theory of the sort discussed here, but to this point it is not nearly as well developed as the commutative theory. All the previous work of which we are aware has been in the context of the noncommutative-ball case (see Example 1.2 above). In this case the Schur-Agler class $`𝒮𝒜_{\mathrm{\Gamma }^{FM}}(𝒰,𝒴)`$ can be identified with the space of contractive multipliers on a Fock space of formal power series in noncommuting indeterminants with norm-square-summable vector coefficients, a noncommutative analogue of the unit ball of analytic Toeplitz operators acting on the classical Hardy space (see and the references there). In particular, Constantinescu and Johnson formulated and obtained a necessary and sufficient condition (in terms of positivity of an associated Pick matrix) for the existence of solutions for an interpolation problem of the form (when translated to our notation) $`F^R(Z_i)=W_i`$ ($`i=1,\mathrm{},N`$) for the class $`𝒮𝒜_{\mathrm{\Gamma }^{FM}}(,)`$. A number of authors (see ) have analyzed noncommutative analogues of the Sarason formulation of interpolation for the noncommutative-ball setting; one approach for these problems is as an application of the Commutant Lifting Theorem developed by Popescu for this setting (see ). A direction for future work is to understand the connections of our approach via evaluation with operator argument with the Sarason formulation and commutant lifting theory. We mention that a very general version of commutant lifting theory (with applications to new sorts of interpolation problems) has recently been worked out by Muhly and Solel . The paper is organized as follows. After the present Introduction, Section 2 derives some consequences of the energy balance relations encoded in the conservative SNMLSs beyond what was derived in which are needed in the sequel. These consequences are then used in Section 3 to derive some necessary conditions for a given pair of operators $`(X_L,T_L)`$ (or $`(T_R,Y_R)`$) to induce a well-defined left (or right) tangential point evaluation with operator argument on a given noncommutative Schur-Agler class $`𝒮𝒜_\mathrm{\Gamma }(𝒰,𝒴)`$. Section 4 then establishes the criterion for existence of solutions in Theorems 1.40 and 1.11. Section 5 establishes a correspondence between solutions of Problem 1.10 and unitary extensions of a certain partially-defined isometry constructed from the data of the problem, while Section 6 then uses the idea of Arov-Grossman (see ) to obtain the linear-fractional parametrization for the set of all solutions of Problem 1.10 as described in Theorem 1.11. Sections 4, 5 and 6 closely parallel the analysis of worked out for the commutative case. The final Section 7 discusses various examples and special cases. ## 2. Conservative structured noncommutative multidimensional linear systems Following we define a structured noncommutative multidimensional linear system (SNMLS) to be a collection $$𝚺=\{\mathrm{\Gamma },,𝒰,𝒴,𝐔\}$$ (2.1) where $`\mathrm{\Gamma }`$ is an admissible graph, $`=\{_p:pP\}`$ is a collection of (separable) Hilbert spaces (called state spaces) indexed by the path-connected components $`p`$ of the graph $`\mathrm{\Gamma }`$, where $`𝒰`$ and $`𝒴`$ are additional (separable) Hilbert spaces (to be interpreted as the input space and the output space respectively) and where $`𝐔`$ is a connection matrix (sometimes also called colligation) of the form $$𝐔=\left[\begin{array}{cc}A& B\\ C& D\end{array}\right]=\left[\begin{array}{cc}[A_{r,s}]& [B_r]\\ [C_s]& D\end{array}\right]:\left[\begin{array}{c}_{sS}_{[s]}\\ 𝒰\end{array}\right]\left[\begin{array}{c}_{rR}_{[r]}\\ 𝒴\end{array}\right]$$ (2.2) In case the connection matrix $`𝐔`$ is unitary, we shall say that $`𝚺`$ is a conservative or unitary SNMLS. Associated with any SNMLS $`𝚺`$ as in (2.1) is the collection of system equations with evolution along the free semigroup $`_E`$ $`𝚺:\{\begin{array}{ccc}\hfill x_{𝐬(e)}(ew)& =& \mathrm{\Sigma }_{sS}A_{𝐫(e),s}x_s(w)+B_{𝐫(e)}u(w)\hfill \\ \hfill x_s^{}(ew)& =& 0\text{ if }s^{}𝐬(e)\hfill \\ \hfill y(w)& =& \mathrm{\Sigma }_{sS}C_sx_s(w)+Du(w)\text{ for }w_E.\hfill \end{array}`$ (2.6) ###### Remark 2.1. Suppose that $$\stackrel{~}{𝚺}=\{\mathrm{\Gamma },\stackrel{~}{},𝒰,𝒴,\stackrel{~}{𝐔}\}$$ (2.7) is another SNMLS with the same structure graph $`\mathrm{\Gamma }`$ and the same input and output spaces as in (2.1) and with the connecting matrix $$\stackrel{~}{𝐔}=\left[\begin{array}{cc}\stackrel{~}{A}& \stackrel{~}{B}\\ \stackrel{~}{C}& \stackrel{~}{D}\end{array}\right]=\left[\begin{array}{cc}[\stackrel{~}{A}_{r,s}]& [\stackrel{~}{B}_r]\\ [\stackrel{~}{C}_s]& \stackrel{~}{D}\end{array}\right]:\left[\begin{array}{c}_{sS}\stackrel{~}{}_{[s]}\\ 𝒰\end{array}\right]\left[\begin{array}{c}_{rR}\stackrel{~}{}_{[r]}\\ 𝒴\end{array}\right].$$ (2.8) The colligations $`𝚺`$ and $`\stackrel{~}{𝚺}`$ are said to be unitarily equivalent if there is a collection $`\mathrm{{\rm Y}}=\{\mathrm{{\rm Y}}_p:pP\}`$ of unitary operators $`\mathrm{{\rm Y}}_p:_p\stackrel{~}{}_p`$ (for each path connected component $`p`$ of $`\mathrm{\Gamma }`$) such that $$\left[\begin{array}{cc}_{rR}\mathrm{{\rm Y}}_{[r]}& 0\\ 0& I_𝒴\end{array}\right]\left[\begin{array}{cc}A& B\\ C& D\end{array}\right]=\left[\begin{array}{cc}\stackrel{~}{A}& \stackrel{~}{B}\\ \stackrel{~}{C}& \stackrel{~}{D}\end{array}\right]\left[\begin{array}{cc}_{sS}\mathrm{{\rm Y}}_{[s]}& 0\\ 0& I_𝒰\end{array}\right].$$ (2.9) It is an easy computation to see that unitarily equivalent colligations have the same transfer functions. It is much less obvious that under certain minimality conditions (structure controllability and observability), the colligations having the same characteristic functions are unitarily equivalent (see \[16, Theorem 7.2\] for the proof). It will be convenient to have the notation $`ps_p`$ for a source-vertex cross-section, i.e., for each path-connected component $`p`$ of $`\mathrm{\Gamma }`$, $`s_p`$ is the assignment of a one particular source vertex in the path-connected component $`p`$. From the structure of the system equations (2.6) and under the assumption that $`𝐔`$ is unitary (or more generally, under the assumption that $`𝐔`$ is contractive), we read off the following properties for system trajectories $`w(u(w),x(w),y(w))`$ satisfying equations (2.6): $`x_s(e_{s,r}w)\text{ is independent of }s\text{ for any given }rR\text{ and }w_E,`$ (2.10) $`{\displaystyle \underset{rR}{}}x_{s_{[r]}}(e_{s_{[r]},r}w)^2x(w)^2u(w)^2y(w)^2,`$ (2.11) $`x_s^{}(ew)=0\text{ if }s^{}𝐬(e).`$ (2.12) We may then compute $`{\displaystyle \underset{eE}{}}x(ew)^2`$ $`={\displaystyle \underset{sS}{}}{\displaystyle \underset{eE}{}}x_s(ew)^2`$ $`={\displaystyle \underset{eE}{}}x_{𝐬(e)}(ew)^2\text{ (by (}\text{2.12}\text{))}`$ $`={\displaystyle \underset{pP}{}}{\displaystyle \underset{sS\text{ and }rR:[s]=[r]}{}}x_s(e_{s,r}w)^2`$ $`={\displaystyle \underset{pP}{}}{\displaystyle \underset{rR:[r]=p}{}}n_{s_p}x_{s_p}(e_{s_p,r}w)^2\text{ (by (}\text{2.10}\text{))}`$ where we have set $`n_{s_p}`$ equal to the number of source vertices $`s`$ in the path-connected component $`p`$ of $`\mathrm{\Gamma }`$. If we now set $`N_S`$ equal to the maximum number of source vertices in any path-connected component of $`\mathrm{\Gamma }`$ $$N_S=\mathrm{max}\{n_{s_p}:pP\},$$ (2.13) then $`{\displaystyle \underset{eE}{}}{\displaystyle \frac{1}{N_S}}x(ew)^2`$ $`={\displaystyle \underset{pP}{}}{\displaystyle \underset{r:[r]=p}{}}{\displaystyle \frac{n_{s_p}}{N_S}}x_{s_p}(e_{s_{p,r}}w)^2`$ $`{\displaystyle \underset{pP}{}}{\displaystyle \underset{r:[r]=p}{}}x_{s_p}(e_{s_p,r}w)^2`$ $`={\displaystyle \underset{rR}{}}x_{s_p}(e_{s_p,r}w)^2`$ $`x(w)^2+u(w)^2y(w)^2\text{ (by (}\text{2.11}\text{)).}`$ Summing over all words $`w`$ of a fixed length $`n`$ and then multiplying by $`N_S^n`$ then gives $`{\displaystyle \underset{w:|w|=n+1}{}}{\displaystyle \frac{1}{N_S^{n+1}}}x(w)^2{\displaystyle \underset{w:|w|=n}{}}{\displaystyle \frac{1}{N_S^n}}x(w)^2`$ $`{\displaystyle \underset{w:|w|=n}{}}{\displaystyle \frac{1}{N_S^n}}u(w)^2{\displaystyle \underset{w:|w|=n}{}}{\displaystyle \frac{1}{N_S^n}}y(w)^2.`$ (2.14) If we now sum over $`n=0,1,\mathrm{},N`$, the left-hand side of (2.14) telescopes and we arrive at $$\underset{w:|w|=N+1}{}\frac{1}{N_S^{|w|}}x(w)^2x(\mathrm{})^2\underset{w:|w|N}{}\frac{1}{N_S^{|w|}}u(w)^2\underset{w:|w|N}{}\frac{1}{N_S^{|w|}}y(w)^2.$$ (2.15) In particular, we get the estimate $$\underset{w:|w|N}{}\frac{1}{N_S^{|w|}}y(w)^2x(\mathrm{})^2+\underset{w:|w|N}{}\frac{1}{N_S^{|w|}}u(w)^2.$$ Letting $`N\mathrm{}`$ then gives $$\underset{w_E}{}\frac{1}{N_S^{|w|}}y(w)^2x(\mathrm{})^2+\underset{w_E}{}\frac{1}{N_S^{|w|}}u(w)^2$$ (2.16) for all system trajectories $`(u,x,y)`$ of the SNMLS $`𝚺`$ as long as the connection matrix $`𝐔`$ satisfies $`𝐔1`$. If $`\{u(w)\}_{w_E}`$ is a $`𝒰`$-valued input string and $`x(\mathrm{})`$ the initial state fed into the system equations to produce a $`𝒴`$-valued output string $`\{y(w)\}_{w_E}`$ and if we introduce the formal $`Z`$-transform of the $`\{u(w)\}_{w_E}`$ and $`\{y(w)\}_{w_E}`$ according to $$\widehat{u}(z)=\underset{w_E}{}u(w)z^w,\widehat{y}(z)=\underset{w_E}{}y(w)z^w,$$ then it follows that $$\widehat{y}(z)=C(IZ_{\mathrm{\Gamma },}(z)A)^1x(\mathrm{})+F_𝚺(z)\widehat{u}(z)$$ (2.17) where $`F_𝚺(z)`$ is the formal noncommutative power series with coefficients in $`(𝒰,𝒴)`$ given by $`F_𝚺(z)`$ $`=`$ $`D+C(IZ_{\mathrm{\Gamma },}(z)A)^1Z_{\mathrm{\Gamma },}(z)B`$ $`=`$ $`F_{\mathrm{}}+{\displaystyle \underset{N=1}{\overset{\mathrm{}}{}}}{\displaystyle \underset{e_1,\mathrm{},e_NE}{}}C_{𝐬(e_N)}A_{𝐫(e_N),𝐬(e_{N1})}\mathrm{}A_{𝐫(e_2),𝐬(e_1)}B_{𝐫(e_1)}z_{e_N}z_{e_{N1}}\mathrm{}z_{e_2}z_{e_1}`$ with $`Z_{\mathrm{\Gamma },}`$ given by (1.19). In particular, if we take the initial state $`x(\mathrm{})`$ equal to $`0`$, we obtain the relation $`\widehat{y}(z)=F_𝚺(z)\widehat{u}(z)`$ between the $`Z`$-transformed input signal $`\widehat{u}(z)`$ and the $`Z`$-transformed output signal $`\widehat{y}(z)`$. We shall call $`F_𝚺(z)`$ the transfer function of the SNMLS $`𝚺`$ (see ). The assertion of Theorem 1.7 then is that a power series $`F`$ belongs to the noncommutative Schur-Agler class $`𝒮𝒜_\mathrm{\Gamma }(𝒰,𝒴)`$ if and only if it is the transfer function of a conservative SNMLS $`𝚺`$ of the form (2.1). ###### Remark 2.2. For future reference, we note that the action of $`F_𝚺(z)`$ on a vector $`u𝒰`$, namely $$F_𝚺(z)=D+C\left(IZ_{\mathrm{\Gamma },}(z)A\right)^1Z_{\mathrm{\Gamma },}(z)B:uy$$ is the result of the feedback connection $$\left[\begin{array}{cc}A& B\\ C& D\end{array}\right]\left[\begin{array}{c}h\\ u\end{array}\right]=\left[\begin{array}{c}h^{}\\ y\end{array}\right],h=Z_{\mathrm{\Gamma },}(z)h^{}$$ where $`h_{sS}\stackrel{~}{}_{[s]}`$ and $`h^{}_{rR}\stackrel{~}{}_{[r]}`$. ###### Remark 2.3. For the special case where $`n_{s_p}=1`$ for each path-connected component and $`𝐔`$ is isometric, it is easily verified that one gets equality in (2.14) and (2.15). Thus, in this case $`N_S=1`$ and (2.16) holds with $`N_S=1`$. All this has already been noted in (see Remark 5.14 there) where such graphs $`G`$ are called row-sum graphs. A particularly nice case of a row sum graph is a Fornasini-Marchesini graph (a row-sum graph with one path-connected component)—see Example 1.2. Then the system and the associated noncommutative function theory have a particularly nice structure—see . ## 3. Admissible interpolation data sets With these preliminaries out of the way, we now turn to the issue of identifying large classes of examples of left-admissible and right-admissible pairs $`(T_L,X_L)`$ and $`(Y_R,T_R)`$ for a general admissible graph $`\mathrm{\Gamma }`$. In particular, we shall see that the class of interpolation problems covered in Problem 1.8 and 1.10 is nonempty. We first note the following relations (stated here without proof) between left and right evaluation with operator argument (1.27) and (1.29) and tensor-product functional calculus (1.18). ###### Proposition 3.1. Assuming that all the functional evaluations below exist, we have the following relations among tensor-product evaluation with operator argument (1.18), left evaluation with operator argument (1.27) and right evaluation with operator argument (1.29). 1. Let $`F(z)=_{v_E}F_vz^v`$ be a formal power series with coefficients in $`(𝒰,𝒴)`$ and let $`T=(T_e:eE)`$ be a tuple of operators on the space $`𝒦`$. Define a new power series $`F^{}(z)`$ by $$F^{}(z)=\underset{v_E}{}F_v^{}z^v\text{if}F(z)=\underset{v_E}{}F_vz^v.$$ Denote by $`(F^{}I_𝒦)(z)`$ the power series $$(F^{}I_𝒦)(z)=\underset{v_E}{}(F_v^{}I_𝒦)z^v.$$ Then $$(F^{}I_𝒦)^L(I_𝒴T)=F(T)=(F^{}I_𝒦)^R(I_𝒰T).$$ (3.1) 2. If $`f(z)=_{v_E}f_vz^v`$ is a formal power series with scalar coefficients (so $`f_v`$ for all $`v_E`$), then $$(f^{}I_𝒦)^L(T)=f(T)=(f^{}I_𝒦)^R(T).$$ (3.2) 3. If $`f(z)=_{v_E}f_vz^v`$ is a formal power series with scalar coefficients as in #2 above and if $`x`$ is a vector in $`𝒦`$, then $$f(T)x=(xf^{})^L(T).$$ (3.3) 4. If $`F(z)=_{v_E}F_vz^v`$ is a formal power series with coefficients $`F_v(𝒰,𝒴)`$, $`\lambda =(\lambda _e)_e`$ is a tuple of complex numbers considered as operators on $``$, and $`X_L(𝒴,𝒦_L)`$ and $`Y_R(𝒦_R,𝒰)`$, then $`(X_LF)^L(\lambda I_{𝒦_L})`$ $`=X_LF(\lambda ),`$ (3.4) $`(FY_R)^R(\lambda I_{𝒦_R})`$ $`=F(\lambda )Y_R.`$ (3.5) ###### Remark 3.2. The left-side of (3.3) is the type of point evaluation used by Rosenblum-Rovnyak to formulate the so-called Nudelman interpolation problem in . Relation (3.3) shows how this type of interpolation condition can be converted to the version of Nudelman interpolation for the classical case used in . An alternative extension of the Rosenblum-Rovnyak Nudelman problem to the formal power series setting is given in . In the sequel we shall have use of only part (4) of Proposition 3.1. By definition, a formal power series $`F(z)=_{v_E}F_vz^v`$ is in the Schur-Agler class if and only if $`F(T)`$ (defined via (1.18)) is a contraction for all $`T_\mathrm{\Gamma }(𝒦)`$. Given operators $`X_L(𝒴,𝒦_L)`$ and $`Y_R(𝒦_R,𝒰)`$ and operator tuples $`T_L(𝒦_L)^{n_E}`$ and $`T_R(𝒦_R)^{n_E}`$ (here we use $`n_E`$ to denote the number of edges $`eE`$ for the admissible graph $`\mathrm{\Gamma }`$), the hope would be that $`(T_L,X_L)`$ would be left admissible as soon as $`T_L_\mathrm{\Gamma }(𝒦_L)`$ and that $`(Y_R,T_R)`$ would be right admissible (with respect to $`𝒮𝒜_\mathrm{\Gamma }(𝒰,𝒴)`$) as soon as $`T_R`$ is in $`_\mathrm{\Gamma }(𝒦_R)`$. As we shall see below, this is indeed correct in some special cases while we obtain only partial results in this direction for the case of a general admissible graph $`\mathrm{\Gamma }`$. We begin with the situation of part (4) in Proposition 3.1. ###### Proposition 3.3. Suppose that $`F(z)=_{v_E}F_vz^v`$ is a formal power series in the class $`𝒮𝒜_\mathrm{\Gamma }(𝒰,𝒴)`$ and suppose that $`\lambda =(\lambda _e)_{eE}`$ is a tuple of complex numbers. Then: 1. Suppose that $`X_L(𝒴,𝒦_L)`$ and that we let $`T_L`$ be the tuple of scalar operators $`T_L=\lambda I_{𝒦_L}=(\lambda _eI_{𝒦_L})_{eE}`$. Then $`(\lambda _eI_{𝒦_L},X_L)`$ is left admissible whenever $`Z_\mathrm{\Gamma }(\lambda )<1`$. 2. Suppose that $`Y_R(𝒦_R,𝒰)`$ and that we let $`T_R`$ be the tuple of scalar operators $`T_R=\lambda I_{𝒦_L}=(\lambda _eI_{𝒦_L})_{eE}`$. Then $`(Y_R,\lambda I_{𝒦_R})`$ is right admissible whenever $`Z_\mathrm{\Gamma }(\lambda )<1`$. 3. Suppose that the hypotheses of parts (1) and (2) hold with $`𝒦_L=𝒴`$ and $`X_L=I_𝒴`$ and with $`𝒦_R=𝒰`$ and $`Y_R=I_𝒰`$. Then $$F^L(\lambda I_𝒴)=F^R(\lambda I_𝒰)$$ (3.6) ###### Proof. This is an immediate consequence of relations (3.4) and (3.5) in Proposition 3.1 and the definitions. ∎ We next explore the function of the scalar-tuple variable $`\lambda =(\lambda _1,\mathrm{},\lambda _{n_E})`$ a little further. To simplify notation, in the statement of the next result we label the edges of the graph $`G`$ by the letters $`1,2,\mathrm{},d`$ where $`d=n_E`$ is the number of edges of $`G`$. Then words in $`_E`$ have the form $`w=i_Ni_{N1}\mathrm{}i_1`$ where each $`i_{\mathrm{}}\{1,\mathrm{},d\}`$. If $`F(z)=_{v_E}F_vz^v`$ is a formal power series with coefficients in $`(𝒰,𝒴)`$, the function $`F^𝐚(\lambda )`$ of the scalar $`d`$-tuple $`(\lambda _1,\mathrm{},\lambda _d)`$ given by either the left-hand side or the right-hand side of (3.6) (under the assumption that the series converges) can be expressed as $`F^𝐚(\lambda )`$ $`={\displaystyle \underset{v_E}{}}F_v(\lambda I_𝒰)^v={\displaystyle \underset{𝐧_+^d}{}}\left[{\displaystyle \underset{v:v𝐚^1(𝐧)}{}}F_v\right]\lambda ^𝐧=:{\displaystyle \underset{𝐧_+^d}{}}F_𝐧^𝐚\lambda ^𝐧`$ where we have introduced the abelianization map $`𝐚:_d_+^d`$ given by $$𝐚(i_N\mathrm{}i_1)=(n_1,\mathrm{},n_d)\text{ if }n_j=\mathrm{\#}\{\mathrm{}:i_{\mathrm{}}=j\}\text{ for }j=1,\mathrm{},d,$$ where $`\lambda ^v=\lambda _{i_N}\mathrm{}\lambda _{i_1}`$ if $`v=i_N\mathrm{}i_1`$ and where $`\lambda ^𝐧=\lambda _1^{n_1}\mathrm{}\lambda _d^{n_d}`$ if $`𝐧=(n_1,\mathrm{},n_d)`$, and where we have set $$F_𝐧^𝐚=\underset{v:v𝐚^1(𝐧)}{}F_v.$$ If $`F𝒮𝒜_\mathrm{\Gamma }(𝒰,𝒴)`$ then necessarily $`F^𝐚`$ is analytic on $`𝒟_{Z_\mathrm{\Gamma }^𝐚}`$ where $`Z_\mathrm{\Gamma }^𝐚(\lambda )`$ is just the abelianization of the structure matrix $`Z_\mathrm{\Gamma }(z)`$ for $`\mathrm{\Gamma }`$. For a general matrix-valued polynomial $`𝐐(\lambda )`$ in the commuting variables $`\lambda =(\lambda _1,\mathrm{},\lambda _d)`$, the associated commutative Schur-Agler class $`𝒮𝒜_𝐐(𝒰,𝒴)`$ was defined in to consist of holomorphic functions $`\lambda F(\lambda )`$ defined on the domain $`𝒟_𝐐:=\{\lambda ^d:𝐐(\lambda )<1\}`$ such that $`F(T)1`$ for any commuting $`d`$-tuple of operators $`(T_1,\mathrm{},T_d)`$ on $`𝒦`$ such that $`𝐐(T_1\mathrm{},T_d)<1`$. For the special case where $`𝐐`$ is taken to be the abelianized structure matrix $`𝐐(\lambda )=Z_\mathrm{\Gamma }^𝐚(\lambda )`$, then we see that the the set of commuting $`d`$-tuples $`T`$ with $`Z_\mathrm{\Gamma }^𝐚(T)<1`$ is just the intersection of $`_\mathrm{\Gamma }(𝒦)`$ with commutative operator tuples. A consequence of Lemma 1 from is that a commuting $`d`$-tuple $`T=(T_1,\mathrm{},T_d)`$ has its Taylor spectrum in the domain $`𝒟_{Z_\mathrm{\Gamma }^𝐚}`$ whenever $`Z_\mathrm{\Gamma }^𝐚(T)<1`$. Moreover, as $`Z_\mathrm{\Gamma }^𝐚`$ is a linear polynomial, the associated domain $`𝒟_{Z_\mathrm{\Gamma }^𝐚}`$ is a logarithmically convex Rinehardt domain, and the functional calculus with operator argument defined via the Taylor functional calculus can equivalently be carried out by using power series centered at the origin (see \[15, Remark 2.2\]). Hence, if $`T_L=(T_{L,j})_{j=1,\mathrm{},d}`$ is a commuting $`d`$-tuple of operators on $`𝒦`$ and $`X_L(𝒴,𝒦_L)`$, then $$(X_LF)^L(T_L)=\underset{v_E}{}T_L^vX_LF_v=\underset{𝐧_+^d}{}T_L^𝐧X_LF_𝐧^𝐚=(X_LF^𝐚)^L(T_L)$$ (3.7) where $`(X_LF^𝐚)^L(T_L)`$ is the functional calculus with commuting operator argument used in . We conclude that: if the formal power series $`F(z)(𝒰,𝒴)z`$ is in the noncommutative Schur-Agler class $`𝒮𝒜_\mathrm{\Gamma }(𝒰,𝒴)`$, then its abelianization $`F^𝐚(\lambda )`$ is in the commutative Schur-Agler class $`𝒮𝒜_{Z_\mathrm{\Gamma }^𝐚}`$ associated with $`𝐐(\lambda ):=Z_\mathrm{\Gamma }^𝐚(\lambda )`$ as defined in . Moreover, we see that the pair $`(X_L,T_L)`$ is left admissible whenever $`T_L=(T_{L,1},\mathrm{},T_{L,d})`$ is a commutative operator-tuple in $`_\mathrm{\Gamma }(𝒦)`$, and then, from the identity (3.7), we see in addition that $$(X_LF)^L(T_L)=(X_LF^a)^L(T_L).$$ More generally, if $`T_L=(T_{L,1},\mathrm{},T_{L,d})`$ is a commuting operator-tuple with Taylor spectrum contained in $`𝒟_{Z_\mathrm{\Gamma }^𝐚}`$, one can use Theorem 2.1 from to see that then $`T_L`$ is similar to a commuting operator-tuple $`T_L^{}`$ satisfying $`Z_\mathrm{\Gamma }^𝐚(T_L^{})<1`$, and hence $`(X_L,T_L)`$ is admissible in this case as well. We have arrived at the following result. ###### Proposition 3.4. Suppose that $`F(z)=_{v_E}F_vz^v`$ is a formal power series in the class $`𝒮𝒜_\mathrm{\Gamma }(𝒰,𝒴)`$. 1. Suppose that $`X_L(𝒴,𝒦_L)`$ and that $`T_L=(T_{L,1},\mathrm{},T_{L,d})`$ is a commutative tuple of operators on $`𝒦_L`$ with Taylor joint spectrum $`\sigma _{\text{Taylor}}(T_L)`$ contained in $`𝒟_{Z_\mathrm{\Gamma }^𝐚}`$. Then the pair $`(T_L,X_L)`$ is left-admissible. In particular, $`(T_L,X_L)`$ is in left-admissible whenever $`T_L`$ is a commutative tuple in $`_\mathrm{\Gamma }(𝒦_L)`$. 2. Suppose that $`Y_R(𝒦_R,𝒰)`$ and that $`T_R=(T_{R,1},\mathrm{},T_{R,d})`$ a commutative tuple of operators on $`𝒦_R`$ with Taylor joint spectrum contained in $`𝒟_{Z_\mathrm{\Gamma }^𝐚}`$. Then the pair $`(Y_R,T_R)`$ is right-admissible. In particular, $`(Y_R,T_R)`$ is right-admissible whenever $`T_R`$ is a commutative tuple in $`_\mathrm{\Gamma }(𝒦_R)`$. ###### Proof. Statement (1) follows from the discussion immediately preceding the statement of the Proposition. A completely parallel argument proves statement (2). ∎ We now give a sufficient condition for left-admissibility for the general case. ###### Proposition 3.5. Suppose that $`\mathrm{\Gamma }`$ is an admissible graph, $`T=\{T_e:eE\}`$ is a tuple of operators on the Hilbert space $`𝒦_L`$ and that $`X_L(𝒴,𝒦_L)`$. Set $`\rho _{_{\mathrm{\Gamma },L}}=1/N_S`$ with $`N_S`$ defined as in (2.13). Then a sufficient condition for the pair $`(T_L,X_L)`$ to be left admissible with respect to the Schur-Agler class $`𝒮𝒜_\mathrm{\Gamma }(𝒰,𝒴)`$ is that $$\underset{v_E}{}\rho _{_{\mathrm{\Gamma },L}}^{|v|}X_L^{}T_L^vk_𝒴^2<\mathrm{}\text{for all }k𝒦_L.$$ (3.8) ###### Proof. Suppose that $`H(z)=_{v_E}H_vz^v`$ is of the form (1.21) in a representation (1.22) for a Schur-Agler class formal power series $`F(z)`$ as in (1.23). Then $$H(z)=C(IZ_{\mathrm{\Gamma },}(z)A)^1$$ where $`F(z)=D+C(IZ_{\mathrm{\Gamma },}(z)A)^1Z_{\mathrm{\Gamma },}(z)B`$ is a unitary realization for $`F(z)`$. Let $`x_{sS}_{[s]}`$. Then from (2.17) which now takes the form $$\widehat{y}(z)=H(z)x(\mathrm{})+F(z)\widehat{u}(z)$$ we see that the coefficients $`H_vx`$ of $`H(z)x`$ amount to the output string $`y(v)=H_vx`$ associated with running the SNMLS $`𝚺=(\mathrm{\Gamma },,𝒰,𝒴,𝐔=\left[\begin{array}{cc}A& B\\ C& D\end{array}\right])`$ with zero input string $`u(v)=0`$ for all $`v_E`$ and with initial state $`x(\mathrm{})=x_0`$. Hence, from (2.16) we see that $$\underset{v_E}{}\rho _{\mathrm{\Gamma },L}^{|v|}H_vx_𝒴^2x_0^2<\mathrm{}.$$ Hence $`{\displaystyle \underset{n=0}{\overset{N}{}}}{\displaystyle \underset{v_E:|v|=n}{}}\left|T_L^v^{}X_LH_vx_0,k_{𝒦_L}\right|={\displaystyle \underset{v_E}{}}\left|\rho _{_{\mathrm{\Gamma },L}}^{|v|/2}H_vx_0,\rho _{_{\mathrm{\Gamma },L}}^{|v|/2}X_L^{}T_L^vk_𝒴\right|`$ $`\left({\displaystyle \underset{v_E}{}}\rho _{\mathrm{\Gamma },L}^{|v|}H_vx_0_𝒴^2\right)^{1/2}\left({\displaystyle \underset{v_E}{}}\rho _{\mathrm{\Gamma },L}^{|v|}X_L^{}T_L^vk_𝒴^2\right)^{1/2}<\mathrm{}.`$ and it follows that $`(T_L,X_L)`$ is left-admissible as wanted. ∎ Given an admissible graph $`\mathrm{\Gamma }`$, we can always associate a new graph $`\mathrm{\Gamma }^{FM}`$ of Fornasini-Marchesini type (as in Example 1.2) by letting $`\mathrm{\Gamma }^{FM}`$ be the admissible graph of Fornasini-Marchesini type having the same edge set $`E`$ as $`\mathrm{\Gamma }`$. This notation appears in the next corollary. ###### Corollary 3.6. Let $`\mathrm{\Gamma }`$ be an admissible graph with associated $`\rho _{_{\mathrm{\Gamma },L}}=1/N_S`$ given by (2.13), let $`T_L=(T_{L,e})_{eE}`$ be a tuple of operators in $`(𝒦_L)`$ and let $`X_L(𝒴,𝒦_L)`$. Then a sufficient condition for $`(T_L,X_L)`$ to be admissible is that $`Z_{\mathrm{\Gamma }^{FM}}(T_L)<\sqrt{\rho _{\mathrm{\Gamma },L}}`$, i.e., that $$\mathrm{row}_{eE}T_{L,e}<\sqrt{\rho _{\mathrm{\Gamma },L}}.$$ In particular, if $`\mathrm{\Gamma }`$ is a Fornasini-Marchesini graph (see Example 1.2), then $`(T_L,X_L)`$ is left admissible with respect to $`𝒮𝒜_\mathrm{\Gamma }(𝒰,𝒴)`$ whenever $`T_L_\mathrm{\Gamma }(𝒦_L)`$. ###### Proof. Set $`r:=Z_{G^{FM}}(T_L)=\left[\begin{array}{cccc}T_{L,e_1}& T_{L,e_2}& \mathrm{}& T_{L,e_d}\end{array}\right]`$. Then the operator $$Z_{G^{FM}}(T)^{}=\left[\begin{array}{c}T_{L,e_1}^{}\\ \mathrm{}\\ T_{L,e_d}^{}\end{array}\right]:𝒦_{eE}𝒦$$ also has norm $`r`$. Hence, for each $`k𝒦_L`$ we have $$\underset{eE}{}T_{L,e}^{}k^2r^2k^2$$ and, more generally, $$\underset{v:|v|=N+1}{}T_L^vk^2r^2\underset{v_E:|v|=N}{}T_L^vk^2.$$ An easy induction argument then gives $$\underset{v_E:|v|=N}{}T_L^vk^2r^{2N}k^2$$ and hence also $$\underset{v_E:|v|=N}{}X_L^{}T_L^Nk^2r^{2N}X_L^{}^2k^2.$$ Hence $$\underset{N=0}{\overset{\mathrm{}}{}}\underset{v_E:|v|=N}{}\rho _{_{\mathrm{\Gamma },L}}^{|v|}X_L^{}T_L^vk^2X_L^{}^2k^2\underset{N=0}{\overset{\mathrm{}}{}}\left(\frac{r}{\sqrt{\rho _{_{\mathrm{\Gamma },L}}}}\right)^{2N}<\mathrm{}$$ if $`r<\sqrt{\rho _{_{\mathrm{\Gamma },L}}}`$. An application of the criterion (3.8) from Proposition 3.5 now completes the proof of Corollary 3.6. ∎ Given an admissible graph $`\mathrm{\Gamma }`$ together with a tuple of operators $`T_R=(T_{R,e})_{eE}`$ of operators on a Hilbert space $`𝒦_R`$ and an operator $`Y_R(𝒦_R,𝒰)`$, there is a sufficient condition for right admissibility of the $`(Y_R,T_R)`$ in the sense of (1.29) dual to condition (3.8) which can be obtained as follows. Note that weak convergence of the series $`_{v_E}F_vY_RT_R^v^{}`$ is equivalent to weak convergence of the adjoint series $$\underset{v_e}{}T_R^vY_R^{}F_v^{}=\underset{v_e}{}T_R^v^{}Y_R^{}F_v^{}^{}$$ which has the same form as (1.27) with $`T_R^{}`$ in place of $`T_L`$, $`Y_R^{}`$ in place of $`X_L`$ and $`F_v^{}^{}`$ in place of $`F_v`$. To apply the results on left-admissibility to get results on right admissibility, we wish to consider $`(T_R^{},Y_R^{})`$ as a left pair acting on the formal power series $$F(z)^{}=\underset{v_E}{}F_v^{}^{}z^v$$ in place of $`F(z)=_{v_E}F_vz^v`$. We know from Theorem 1.7 that the formal power series $`F(z)=_{v_E}F_vz^v`$ is in the Schur-Agler class $`𝒮𝒜_\mathrm{\Gamma }(𝒰,𝒴)`$ if and only if $`F(z)`$ has a representation (1.20) with $`U=\left[\begin{array}{cc}A& B\\ C& D\end{array}\right]`$ unitary. If $`F(z)`$ has the form (1.20), then we compute $`F(z)^{}`$ $`={\displaystyle \underset{v_E}{}}F_v^{}z^v^{}`$ $`=D^{}+B^{}Z_{\mathrm{\Gamma },}(z)^{}(IA^{}Z_{\mathrm{\Gamma },}(z)^{})^1C^{}`$ $`=D^{}+B^{}(IZ_{\mathrm{\Gamma },}(z)^{}A^{})^1Z_{\mathrm{\Gamma },}(z)^{}C^{}.`$ This suggests that, given a SNMLS $`𝚺=(\mathrm{\Gamma },,𝒰,𝒴,𝐔)`$ as defined in (2.1), we define a dual SNMLS $`𝚺^{}=(\mathrm{\Gamma }^{},,𝒴,𝒰,𝐔^{})`$ where 1. the admissible graph $`\mathrm{\Gamma }^{}`$ for $`𝚺^{}`$ is the same graph as the admissible graph $`\mathrm{\Gamma }`$, but with the source vertices for $`\mathrm{\Gamma }`$ taken to be the range vertices for $`\mathrm{\Gamma }^{}`$ and with the range vertices for $`\mathrm{\Gamma }`$ taken to be the source vertices for $`\mathrm{\Gamma }^{}`$; thus the set of path-components remains unchanged: $`P^{}=P`$, and 2. the connection matrix $`𝐔^{}`$ for $`𝚺^{}`$ is simply the adjoint $$𝐔^{}=𝐔^{}=\left[\begin{array}{cc}A^{}& C^{}\\ B^{}& D^{}\end{array}\right]:\left[\begin{array}{c}_{rR}_{[r]}\\ 𝒴\end{array}\right]\left[\begin{array}{c}_{sS}_{[s]}\\ 𝒰\end{array}\right]$$ of the connection matrix $`U`$ for $`𝚺`$. Then it is easily checked: if $`F(z)`$ is the transfer function of the SNMLS $`𝚺`$, then $`F(z)^{}`$ is the transfer function of the SNMLS $`𝚺^{}`$. Moreover $`𝚺`$ is conservative (i.e., $`𝐔`$ is unitary) if and only if $`𝚺^{}`$ is conservative (i.e., $`𝐔^{}=𝐔^{}`$ is unitary). By the equivalence (1) $``$ (2) in Theorem 1.7, we conclude that: the formal power series $`F(z)=_{v_E}F_vz^v`$ is in the Schur-Agler class $`𝒮𝒜_\mathrm{\Gamma }(𝒰,𝒴)`$ if and only if its adjoint $`F(z)^{}=_{v_E}F_v^{}^{}z^v`$ is in the Schur-Agler class $`𝒮𝒜_\mathrm{\Gamma }^{}(𝒴,𝒰)`$, where $`\mathrm{\Gamma }^{}`$ is the reflection of $`\mathrm{\Gamma }`$ induced by interchanging source vertices with range vertices. A consequence of this analysis is that we have the following analogues of Proposition 3.5 and Corollary 3.6. We leave the details of the proof to the reader. In the statement of the theorem we use the notation $$n_R=\mathrm{max}\{n_{r_p}:pP\}$$ (3.9) where $`n_{r_p}`$ is the number of range vertices in component $`P`$ of the graph $`G`$. ###### Proposition 3.7. Let $`\mathrm{\Gamma }`$ be an admissible graph with associated constant $`\rho _{\mathrm{\Gamma },R}:=1/n_R`$ with $`n_R`$ as in (3.9), let $`T_R=(T_{R,e})_{eE}`$ be a tuple of operators acting on a Hilbert space $`𝒦_R`$ and let $`Y_R(𝒦_R,𝒰)`$. Then a sufficient condition for the pair $`(Y_R,T_R)`$ to be right admissible in the sense of (1.29) is that $$\underset{v_E}{}\rho _{\mathrm{\Gamma },R}^{|v|}Y_RT_R^vk^2<\mathrm{}\text{for all }k𝒦_R.$$ (3.10) For the statement of the following corollary, we use the notation $`\mathrm{\Gamma }^{FM^{}}`$ to denote the dual of the Fornasini-Marchesini graph $`\mathrm{\Gamma }^{FM}`$ associated with $`\mathrm{\Gamma }`$; thus $`\mathrm{\Gamma }^{FM^{}}`$ has a single range vertex $`\{r_0\}`$, the same edge set $`E`$ as does $`\mathrm{\Gamma }`$ and the source-vertex set taken also equal to $`E`$ and with each edge $`e`$ considered to have source itself $`e`$ and range $`r_0`$. The associated structure matrix $`Z_\mathrm{\Gamma }(z)`$ is then a column $$Z_{\mathrm{\Gamma }^{FM^{}}}(z)=\left[\begin{array}{c}z_{e_1}\\ \mathrm{}\\ z_{e_d}\end{array}\right]$$ where $`d=n_E`$ is the number of edges. ###### Corollary 3.8. Let $`\mathrm{\Gamma }`$ be an admissible graph with associated $`\rho _{\mathrm{\Gamma },R}=1/N_R`$ given by (3.9), let $`T_R=(T_{R,e})_{eE}`$ be a tuple of operators in $`(𝒦_R)`$ and let $`X_R(𝒦_R,𝒰)`$. Then a sufficient condition for $`(Y_R,T_R)`$ to be right-admissible is that $`Z_{G^{FM}}(T_R)<\sqrt{\rho _{\mathrm{\Gamma },R}}`$. In particular, if $`\mathrm{\Gamma }=\mathrm{\Gamma }^{FM}`$ is itself the reflection of a Fornasini-Marchesini graph, then $`(Y_R,T_R)`$ is right admissible whenever $`T_R_\mathrm{\Gamma }(𝒦)`$. ## 4. The solvability criterion In this section we prove the necessity part of Theorem 1.40. First we need to note the following elementary properties of evaluations (1.27) and (1.29). ###### Lemma 4.1. Let $`T=\{T_e:eE\}`$ and $`T^{}=\{T_e^{}:eE\}`$ be tuples of bounded linear operators acting on Hilbert spaces $`𝒦`$ and $`𝒦^{}`$, respectively. 1. For every constant function $`W(z)W(𝒦^{},𝒦)`$, $$\left(W\right)^L(T)=\left(W\right)^R(T^{})=W.$$ (4.1) 2. For every $`F(𝒰,𝒦)z`$, and $`\stackrel{~}{F}(𝒦^{},𝒴)z`$, $`W(𝒰^{},𝒰)`$ and $`\stackrel{~}{W}(𝒴,𝒴^{})`$. $$\left(FW\right)^L(T)=F^L(T)W\text{and}\left(\stackrel{~}{W}\stackrel{~}{F}\right)^R(T^{})=\stackrel{~}{W}\stackrel{~}{F}^R(T^{})$$ (4.2) whenever $`F^L(T)`$ and $`\stackrel{~}{F}^R(T^{})`$ are defined. 3. For every $`F`$ and $`\stackrel{~}{F}`$ as in part (2) and every $`eE`$, $$\left(F(z)z_e\right)^L(T)=T_eF^L(T)\text{and}\left(z_e\stackrel{~}{F}(z)\right)^R(T^{})=\stackrel{~}{F}^R(T^{})T_e^{}$$ (4.3) whenever $`F^L(T)`$ and $`\stackrel{~}{F}^R(T^{})`$ are defined. 4. For every choice of $`F(𝒰,𝒦)z`$ and of $`\stackrel{~}{F}(𝒰^{},𝒰)z`$ $$\left(F\stackrel{~}{F}\right)^L(T)=(F^L(T)\stackrel{~}{F})^L(T)$$ (4.4) whenever $`F^L(T)`$ and $`(F^L(T)\stackrel{~}{F})^L(T)`$ are defined. 5. For every choice of $`F(𝒴^{},𝒴)z`$ and of $`\stackrel{~}{F}(𝒦^{},𝒴)z`$, $$\left(F\stackrel{~}{F}\right)^R(T^{})=(F\stackrel{~}{F}^R(T^{}))^R(T^{})$$ (4.5) whenever $`\stackrel{~}{F}^R(T^{})`$ and $`(F\stackrel{~}{F}^R(T^{}))^R(T^{})`$ are defined. Proof: The two first statements follow immediately from definitions (1.27) and (1.29). To prove (4.4), take $`F`$ and $`\stackrel{~}{F}`$ in the form $$F(z)=\underset{v_E}{}F_vz^v,\stackrel{~}{F}(z)=\underset{v_E}{}\stackrel{~}{F}_vz^v.$$ Then $$F(z)\stackrel{~}{F}(z)=\underset{v_E}{}\left(\underset{uw=v}{}F_u\stackrel{~}{F}_w\right)z^v$$ and therefore, according to (1.27), $$\left(F\stackrel{~}{F}\right)^L(T)=\underset{v_E}{}T^v^{}\left(\underset{uw=v}{}F_u\stackrel{~}{F}_w\right).$$ (4.6) On the other hand, again by (1.27), $`(F^L(T)\stackrel{~}{F})^L(T)`$ $`=`$ $`{\displaystyle \underset{w_E}{}}T^v^{}F^L(T)\stackrel{~}{F}_w`$ $`=`$ $`{\displaystyle \underset{w_E}{}}T^w^{}\left({\displaystyle \underset{u_E}{}}T^u^{}F_u\right)\stackrel{~}{F}_w`$ $`=`$ $`{\displaystyle \underset{w,u_E}{}}T^{(uw)^{}}F_u\stackrel{~}{F}_w`$ $`=`$ $`{\displaystyle \underset{v_E}{}}T^v^{}\left({\displaystyle \underset{uw=v}{}}F_u\stackrel{~}{F}_w\right).`$ Comparison of the last equality with (4.6) gives (4.4). Equality (4.5) is obtained in much the same way. The first equality in (4.3) follows from (4.4) for the special case of $`\stackrel{~}{F}(z)=z_eI_𝒰`$. The second equality in (4.3) follows from (4.5) for the special case of $`F(z)=z_eI_𝒴`$.∎ Proof of the necessity part in Theorem 1.40 and 1.11: Let $`F`$ belong to $`𝒮𝒜_\mathrm{\Gamma }(𝒰,𝒴)`$ and suppose that $`F`$ is a solution of Problem 1.8. Choose formal power series $`H`$ and $`G`$ of the form (1.21) and (1.23) so that the representations (1.22), (1.24), (1.25) hold. Use (1.44)–(1.46) to define operators $`\mathrm{\Psi }_{s,s^{}}`$, $`\mathrm{\Lambda }_{s,r}`$ and $`\mathrm{\Phi }_{r,r^{}}`$ for $`s,s^{}S`$ and $`r,r^{}R`$. Then use equations (1.41)–(1.43) to define the block operator matrix $`𝕂_p`$. If $`F`$ is assumed to be a solution of Problem 1.10 then we are given $`𝕂_p`$ via (1.41)–(1.43) where $`\mathrm{\Psi }_{s,s^{}}`$, $`\mathrm{\Lambda }_{s,r}`$ and $`\varphi _{r,r^{}}`$ are part of the interpolation data and (1.44)–(1.46) hold as part of the interpolation conditions for some choice of $`H`$ and $`G`$ associated with the representations (1.22), (1.24), (1.25) for $`F`$. In any case, the conditions (1.44)–(1.46) hold and imply that $`𝕂_p`$ can be represented as $$𝕂_p=\left[\begin{array}{c}𝕋_{p,L}^{}\\ 𝕋_{p,R}^{}\end{array}\right]\left[\begin{array}{cc}𝕋_{p,L}& 𝕋_{p,R}\end{array}\right]$$ (4.7) where the operators $`𝕋_{p,L}`$ and $`𝕋_{p,R}`$ are given by $`𝕋_{p,L}`$ $`=`$ $`\mathrm{Row}_{sS:[s]=p}\left[(X_LH_s)^L(T_L)\right]^{}:{\displaystyle \underset{sS:[s]=p}{}}𝒦_L_p,`$ (4.8) $`𝕋_{p,R}`$ $`=`$ $`\mathrm{Row}_{rR:[r]=p}(G_rY_R)^R(T_R):{\displaystyle \underset{rR:[r]=p}{}}𝒦_R_p.`$ (4.9) Comparing (4.7) with (1.41) we see that $$𝕂_{p,L}=𝕋_{p,L}^{}𝕋_{p,L},𝕂_{p,R}=𝕋_{p,R}^{}𝕋_{p,R},𝕂_{p,LR}=𝕋_{p,L}^{}𝕋_{p,R}.$$ (4.10) It follows from (4.7) that $`𝕂_p0`$ and thus, it remains to show that these operators satisfy the Stein identity (1.39). To this end, note that by (1.19) and (1.21), $$H(z)Z_{\mathrm{\Gamma },}(z)=\mathrm{Row}_{rR}\underset{sS:[s]=[r]}{}H_s(z)z_{e_{s,r}}$$ and therefore, by the first equality in (4.3), $$\left(X_LHZ_{\mathrm{\Gamma },}\right)^L(T_L)=\mathrm{Row}_{rR}\underset{sS:[s]=[r]}{}T_{L,e_{s,r}}\left(X_LH_s\right)^L(T_L)$$ which can be written in terms of (1.35) and (4.8) as $$\left(X_LHZ_{\mathrm{\Gamma },}\right)^L(T_L)=\mathrm{Row}_{rR}\stackrel{~}{N}_r(T_L)^{}𝕋_{[r],L}^{}.$$ (4.11) Note also that according to decompositions (1.21) and (1.33), $$\left(X_LH\right)^L(T_L)=\mathrm{Row}_{sS}E_{L,s}^{}𝕋_{[s],L}^{}.$$ (4.12) Similarly, by (1.19) and (1.23), $$Z_{\mathrm{\Gamma },}(z)G(z)=\mathrm{Col}_{sS}\underset{rR:[r]=[s]}{}z_{e_{s,r}}G_r(z)$$ and therefore, by the second equality in (4.3), $$\left(Z_{\mathrm{\Gamma },}GY_R\right)^R(T_R)=\mathrm{Col}_{sS}\underset{rR:[r]=[s]}{}\left(G_rY_R\right)^R(T_R)T_{R,e_{s,r}},$$ which can be written in terms of (1.36) and (4.9) as $$\left(Z_{\mathrm{\Gamma },}GY_R\right)^R(T_R)=\mathrm{Col}_{sS}𝕋_{[s],R}\stackrel{~}{M}_s(T_R).$$ (4.13) Finally, by decompositions (1.23) and (1.34), $$\left(GY_R\right)^R(T_R)=\mathrm{Col}_{rR}𝕋_{[r],R}E_{R,r}.$$ (4.14) Substituting the partitionings (1.37), (1.38), (1.40) and (1.41) into (1.39) we conclude that (1.39) is equivalent to the following three equalities: $`{\displaystyle \underset{sS}{}}E_{L,s}^{}𝕂_{[s],L}E_{L,s}{\displaystyle \underset{rR}{}}\stackrel{~}{N}_r(T_L)^{}𝕂_{[r],L}\stackrel{~}{N}_r(T_L)`$ $`=`$ $`X_LX_L^{}Y_LY_L^{},`$ (4.15) $`{\displaystyle \underset{sS}{}}E_{L,s}^{}𝕂_{[s],LR}\stackrel{~}{M}_s(T_R){\displaystyle \underset{rR}{}}\stackrel{~}{N}_r(T_L)^{}𝕂_{[r],LR}E_{R,r}`$ $`=`$ $`X_LX_RY_LY_R,`$ (4.16) $`{\displaystyle \underset{sS}{}}\stackrel{~}{M}_s(T_R)^{}𝕂_{[s],R}\stackrel{~}{M}_s(T_R){\displaystyle \underset{rR}{}}E_{R,r}^{}𝕂_{[r],R}E_{R,r}`$ $`=`$ $`X_R^{}X_RY_R^{}Y_R.`$ (4.17) To check (4.15) we consider the equality $$X_LX_L^{}X_LF(z)F(z^{})^{}X_L^{}=X_LH(z)\left(IZ_{\mathrm{\Gamma },}(z)Z_{\mathrm{\Gamma },}(z^{})^{}\right)H(z^{})^{}X_L^{}$$ (4.18) which is an immediate corollary of (1.22). We may consider each side of (4.18) as a formal power series in $`z^{}`$ with coefficients equal to formal power series in $`z`$, i.e., we have a natural identification $$(𝒦_L)z,z^{}\left((𝒦_L)z\right)z^{}.$$ We then apply the left evaluation map (applied to formal power series in the variable $`z`$) to each coefficient of the resulting formal power series in the variable $`z^{}`$. The result amounts to applying left evaluation to both sides of (4.18) in the variable $`z`$ with the formal variable $`z^{}`$ considered as fixed. Making use of properties (4.1), (4.2) and of relation (4.11) and taking into account the first interpolation condition in (1.32), we get $`X_LX_L^{}Y_LF(z^{})^{}X_L^{}`$ $`=`$ $`\left(X_LH\right)^L(T_L)H(z^{})^{}X_L^{}`$ $`\left(X_LHZ_{\mathrm{\Gamma },}\right)^L(T_L)Z_{\mathrm{\Gamma },}(z^{})^{}H(z^{})^{}X_L^{}.`$ This equality holds as an identity in $`(𝒦_L)z^{}`$. Taking adjoints and replacing $`z^{}`$ by $`z`$, we get $`X_LX_L^{}X_LF(z)Y_L^{}`$ $`=`$ $`X_LH(z)\left(\left(X_LH\right)^L(T_L)\right)^{}`$ $`X_LH(z)Z_{\mathrm{\Gamma },}(z)\left(\left(X_LHZ_{\mathrm{\Gamma },}\right)^L(T_L)\right)^{}.`$ Applying again the left evaluation to the latter equality we get $`X_LX_L^{}Y_LY_L^{}`$ $`=`$ $`\left(X_LH\right)^L(T_L)\left(\left(X_LH\right)^L(T_L)\right)^{}`$ $`\left(X_LHZ_{\mathrm{\Gamma },}\right)^L(T_L)\left(\left(X_LHZ_{\mathrm{\Gamma },}\right)^L(T_L)\right)^{}.`$ Substituting (4.11) and (4.12) into the right hand side expression we come to $$X_LX_L^{}Y_LY_L^{}=\underset{sS}{}E_{L,s}^{}𝕋_{[s],L}^{}𝕋_{[s],L}E_{L,s}\underset{rR}{}\stackrel{~}{N}_r(T_L)^{}𝕋_{[s],L}^{}𝕋_{[s],L}\stackrel{~}{N}_r(T_L)$$ which is equivalent to (4.15), since $$𝕋_{[s],L}^{}𝕋_{[s],L}=𝕂_{[s],L}\text{and}𝕋_{[r],L}^{}𝕋_{[r],L}=𝕂_{[r],L}.$$ To prove (4.16) we start with equality $$X_LF(z)Y_RX_LF(z^{})Y_R=X_LH(z)\left(Z_{\mathrm{\Gamma },}(z)Z_{\mathrm{\Gamma },}(z^{})\right)G(z^{})Y_R$$ which is a consequence of (1.25). We apply the left evaluation in the $`z`$ variable: by the first interpolation condition in (1.32) we have $$Y_LY_RX_LF(z^{})Y_R=\left(X_LHZ_{\mathrm{\Gamma },}\right)^L(T_L)G(z^{})Y_R\left(X_LH\right)^L(T_L)Z_{\mathrm{\Gamma },}(z^{})G(z^{})Y_R.$$ The last identity equality holds true as an identity between formal power series in the variable $`z^{}`$; we then apply the right evaluation (1.29) to both sides. In view of the second interpolation condition in (1.32) and of properties (4.1), (4.2), we obtain $`Y_LY_RX_LX_R`$ $`=`$ $`\left(X_LHZ_{\mathrm{\Gamma },}\right)^L(T_L)\left(GY_R\right)^R(T_R)`$ $`\left(X_LH\right)^L(T_L)\left(Z_{\mathrm{\Gamma },}GY_R\right)^R(T_R).`$ Substituting equalities (4.11), (4.12), (4.13) and (4.14) into the right-hand side expression in the last equality we come to $$X_LX_RY_LY_R=\underset{sS}{}E_{L,s}^{}𝕋_{[s],L}^{}𝕋_{[s],R}\stackrel{~}{M}_s(T_R)\underset{rR}{}\stackrel{~}{N}_r(T_L)^{}𝕋_{[r],L}^{}𝕋_{[r],R}E_{R,r}$$ which is equivalent to (4.16), since $$𝕋_{[s],L}^{}𝕋_{[s],R}=𝕂_{[s],LR}\text{and}𝕋_{[r],L}^{}𝕋_{[r],R}=𝕂_{[r],LR},$$ by (4.10). The proof of (4.17) is quite similar: we start with the equality $$Y_R^{}Y_RY_R^{}F(z)^{}F(z^{})Y_R=Y_R^{}G(z)^{}\left(IZ_{\mathrm{\Gamma },}(z)^{}Z_{\mathrm{\Gamma },}(z^{})\right)G(z^{})Y_R$$ (which follows from (1.24)) and apply the right evaluation in the $`z^{}`$ variable. Then we take adjoints in the resulting formal power series identity (in the variable $`z`$) and apply again the right evaluation map. The obtained equality together with relations (4.13) and (4.14) leads to (4.17). This completes the proof of necessity in both Theorem 1.40 and Theorem 1.11. ## 5. Solutions to the interpolation problem and unitary extensions In this Section we shall show that there is a correspondence between solutions to Problem 1.10 and unitary extensions of a partially defined isometry determined by the problem data set $`𝒟`$. From now on we assume that we are given an interpolation data set $`𝒟`$ as in (1.47) and that the necessary conditions for Problem 1.10 to have a solution are in force: the operators $`𝕂_p`$ defined in (1.41), (1.42) are each positive semidefinite on the space $$_p^{}=\left(\underset{sS:[s]=p}{}𝒦_L\right)\left(\underset{rR:[r]=p}{}𝒦_R\right)$$ (5.1) and satisfy the Stein identity (1.39) which we write now as $$\underset{sS}{}M_s^{}𝕂_{[s]}M_s+Y^{}Y=\underset{rR}{}N_r^{}𝕂_{[r]}N_r+X^{}X.$$ (5.2) For every $`pP`$, we introduce the equivalence $`\stackrel{p}{}`$ on $`_p^{}`$ by $$h_1\stackrel{p}{}h_2\text{if and only if}𝕂_p(h_1h_2),y__p^{}=0\text{for all}y_p^{},$$ denote $`[h]_p`$ the equivalence class of $`h`$ with respect to the above equivalence and endow the linear space of equivalence classes with the inner product $$[h]_p,[y]_p=𝕂_ph,y__p^{}.$$ (5.3) We get a prehilbert space whose completion is $`\widehat{}_p`$. It is readily seen from definitions (1.37), (1.38) of operators $`M_s`$ and $`N_r`$ that $`M_sf`$ and $`N_rf`$ belong to $`_{[s]}^{}`$ and $`_{[r]}^{}`$, respectively, for every choice of $`f𝒦_L𝒦_R`$. Furthermore, identity (5.2) can be written as $$\underset{sS}{}[M_sf]_{[s]}_{\widehat{}_{[s]}}^2+Yf_𝒰^2=\underset{rR}{}[N_rf]_{[r]}_{\widehat{}_{[r]}}^2+Xf_𝒴^2,$$ holding for every choice of $`f𝒦_L𝒦_R`$. Therefore the linear map defined by the rule $$𝐕:\left[\begin{array}{c}\mathrm{Col}_{sS}[M_sf]_{[s]}\\ Yf\end{array}\right]\left[\begin{array}{c}\mathrm{Col}_{rR}[N_rf]_{[r]}\\ Xf\end{array}\right]$$ (5.4) extends by linearity to define an isometry from $$𝒟_𝐕=\mathrm{Clos}\left\{\left[\begin{array}{c}\mathrm{Col}_{sS}[M_sf]_{[s]}\\ Yf\end{array}\right],f𝒦_L𝒦_R\right\}\left[\begin{array}{c}_{sS}\widehat{}_{[s]}\\ 𝒰\end{array}\right]$$ (5.5) onto $$_𝐕=\mathrm{Clos}\left\{\left[\begin{array}{c}\mathrm{Col}_{rR}[N_rf]_{[r]}\\ Xf\end{array}\right],f𝒦_L𝒦_R\right\}\left[\begin{array}{c}_{rR}\widehat{}_{[r]}\\ 𝒴\end{array}\right].$$ (5.6) The next two lemmas establish a correspondence between solutions $`F`$ to Problem 1.10 and unitary extensions of the partially defined isometry $`𝐕`$ given in (5.4). ###### Lemma 5.1. Any solution $`F`$ to Problem 1.10 is a characteristic function of a unitary colligation $$\stackrel{~}{𝚺}=\{\mathrm{\Gamma },\stackrel{~}{},𝒰,𝒴,\stackrel{~}{𝐔}\}$$ (5.7) with the state space $$\stackrel{~}{}=\widehat{}^{}:=\{\stackrel{~}{}_p=\widehat{}_p_p^{}:pP\}$$ and the connecting operator $$\stackrel{~}{𝐔}=\left[\begin{array}{cc}\stackrel{~}{A}& \stackrel{~}{B}\\ \stackrel{~}{C}& \stackrel{~}{D}\end{array}\right]:\left[\begin{array}{c}_{sS}(\widehat{}_{[s]}\stackrel{~}{}_{[s]})\\ 𝒰\end{array}\right]\left[\begin{array}{c}_{rR}(\widehat{}_{[r]}\stackrel{~}{}_{[r]})\\ 𝒴\end{array}\right]$$ (5.8) being an extension of the isometry $`𝐕`$ given in (5.4). Proof: Let $`F`$ be a solution to Problem 1.10. In particular, $`F`$ belongs to the noncommutative Schur-Agler class $`𝒮𝒜_\mathrm{\Gamma }(𝒰,𝒴)`$ and, by Theorem 1.7, it is the characteristic function of some unitary colligation $`𝚺`$ of the form (2.1). In other words, $`F`$ admits a unitary realization (1.20) with the state space $`=\{_p:pP\}`$ and representations (1.22), (1.24), (1.25) hold for power series $`H`$ and $`G`$ defined via (1.26) and decomposed as in (1.21) and (1.23). These series lead to the following two representations $$F(z)=D+H(z)Z_{\mathrm{\Gamma },}(z)B=D+CZ_{\mathrm{\Gamma },}(z)G(z),$$ (5.9) of $`F`$, each of which is equivalent to (1.20). The interpolation conditions (1.32) and (1.44)–(1.46) which hold for $`F`$ by assumption force certain restrictions on the connecting operator $`𝐔=\left[\begin{array}{cc}A& B\\ C& D\end{array}\right]`$. Substituting (5.9) into (1.32) we get equalities $$\left(X_LD+X_LHZ_{\mathrm{\Gamma },}B\right)^L(T_L)=Y_L$$ and $$\left(DY_R+CZ_{\mathrm{\Gamma },}GY_R\right)^R(T_R)=X_R$$ which are equivalent, due to properties (4.1), (4.2), to $$X_LD+\left(X_LHZ_{\mathrm{\Gamma },}\right)^L(T_L)B=Y_L$$ (5.10) and $$DY_R+C\left(Z_{\mathrm{\Gamma },}GY_R\right)^R(T_R)=X_R,$$ (5.11) respectively. It also follows from (1.26) that $$C+H(z)Z_{\mathrm{\Gamma },}(z)A=H(z),B+AZ_{\mathrm{\Gamma },}(z)G(z)=G(z)$$ and therefore, that $$X_LC+\left(X_LHZ_{\mathrm{\Gamma },}\right)^L(T_L)A=\left(X_LH\right)^L(T_L)$$ (5.12) and $$BY_R+A\left(Z_{\mathrm{\Gamma },}GY_R\right)^R(T_R)=\left(GY_R\right)^R(T_R).$$ (5.13) The equalities (5.10) and (5.12) can be written in matrix form as $$\left[\begin{array}{cc}\left(X_LHZ_{\mathrm{\Gamma },}\right)^L(T_L)& X_L\end{array}\right]\left[\begin{array}{cc}A& B\\ C& D\end{array}\right]=\left[\begin{array}{cc}\left(X_LH\right)^L(T_L)& Y_L\end{array}\right],$$ (5.14) whereas the equalities (5.11) and (5.13) are equivalent to $$\left[\begin{array}{cc}A& B\\ C& D\end{array}\right]\left[\begin{array}{c}\left(Z_{\mathrm{\Gamma },}GY_R\right)^R(T_R)\\ Y_R\end{array}\right]=\left[\begin{array}{c}\left(GY_R\right)^R(T_R)\\ X_R\end{array}\right].$$ (5.15) Since the operator $`\left[\begin{array}{cc}A& B\\ C& D\end{array}\right]`$ is unitary, we conclude from (5.14) that $$\left[\begin{array}{cc}A& B\\ C& D\end{array}\right]\left[\begin{array}{c}\left[\left(X_LH\right)^L(T_L)\right]^{}\\ Y_L^{}\end{array}\right]=\left[\begin{array}{c}\left[\left(X_LHZ_{\mathrm{\Gamma },}\right)^L(T_L)\right]^{}\\ X_L^{}\end{array}\right].$$ (5.16) Combining (5.15) and (5.16) we conclude that for every choice of $`f𝒦_L𝒦_R`$, $`\left[\begin{array}{cc}A& B\\ C& D\end{array}\right]\left[\begin{array}{cc}\left[\left(X_LH\right)^L(T_L)\right]^{}& \left(Z_{\mathrm{\Gamma },}GY_R\right)^R(T_R)\\ Y_L^{}& Y_R\end{array}\right]f`$ (5.21) $`=\left[\begin{array}{cc}\left[\left(X_LHZ_{\mathrm{\Gamma },}\right)^L(T_L)\right]^{}& \left(GY_R\right)^R(T_R)\\ X_L^{}& X_R\end{array}\right]f.`$ (5.24) Let $`𝕋_{p,L}`$ and $`𝕋_{p,R}`$ be the operators given by (4.8) and (4.9), respectively, and let $$𝕋_p:=\left[\begin{array}{cc}𝕋_{p,L}& 𝕋_{p,R}\end{array}\right]:_p^{}_p.$$ (5.25) Now we use the interpolation conditions (1.44)–(1.46), which provide the factorization (4.7) of the operator $`𝕂_p`$. Thus, $$𝕂_p=𝕋_p^{}𝕋_p$$ and $$[h]_p,[y]_p_{\widehat{}_p}=𝕂_ph,y__p^{}=𝕋_ph,𝕋_py__p$$ for every $`h,y_p^{}`$. Therefore, the linear transformation $`U_p`$ defined by the rule $$U_p:𝕋_ph[h]_p(h_p^{})$$ (5.26) can be extended to the unitary map (which still is denoted by $`U_p`$) from $`\overline{\mathrm{Ran}𝕋_p}`$ onto $`\widehat{}_p`$. Noticing that $`\overline{\mathrm{Ran}𝕋_p}`$ is a subspace of $`_p`$ and setting $$𝒩_p:=_p\overline{\mathrm{Ran}𝕋_p}\mathrm{and}\stackrel{~}{}_p:=\widehat{}_p𝒩_p,$$ we define the unitary map $`\stackrel{~}{U}_p:_p\stackrel{~}{}_p`$ by the rule $$\stackrel{~}{U}_pg=\{\begin{array}{ccc}U_pg& \mathrm{for}\hfill & g\overline{\mathrm{Ran}𝕋_p},\hfill \\ g& \mathrm{for}\hfill & g𝒩_p.\hfill \end{array}$$ (5.27) Introducing the operators $$\stackrel{~}{A}=\left[\stackrel{~}{U}_{[r]}A_{r,s}\stackrel{~}{U}_{[s]}^{}\right]_{rR,sS},\stackrel{~}{B}=\mathrm{Col}_{rR}\stackrel{~}{U}_{[r]}B_r,\stackrel{~}{C}=\mathrm{Row}_{sS}C_s\stackrel{~}{U}_{[s]}^{},\stackrel{~}{D}=D,$$ (5.28) we construct the colligation $`\stackrel{~}{𝚺}`$ via (2.7) and (5.8). By definition, $`\stackrel{~}{𝚺}`$ is unitarily equivalent to the initial colligation $`𝚺`$ defined in (2.1). By Remark 2.1, $`\stackrel{~}{𝚺}`$ has the same characteristic function as $`𝚺`$, that is, $`F(z)`$. It remains to check that the connecting operator of $`\stackrel{~}{𝚺}`$ is an extension of $`𝐕`$, that is $$\left[\begin{array}{cc}\stackrel{~}{A}& \stackrel{~}{B}\\ \stackrel{~}{C}& \stackrel{~}{D}\end{array}\right]\left[\begin{array}{c}\mathrm{Col}_{sS}[M_sf]_{[s]}\\ Yf\end{array}\right]=\left[\begin{array}{c}\mathrm{Col}_{rR}[N_rf]_{[r]}\\ Xf\end{array}\right]\text{ for every }f𝒦_L𝒦_R.$$ (5.29) To this end, note that by (5.26), (5.27) and block partitionings (1.37) and (5.25) of $`M_s`$ and $`𝕋`$, it holds that $$\stackrel{~}{U}_{[s]}^{}[M_sf]_{[s]}=𝕋_{[s]}(M_sf)=\left[\begin{array}{cc}𝕋_{[s,],L}E_{L,s}& 𝕋_{[s],R}\stackrel{~}{M}_s(T_R)\end{array}\right]f$$ for every $`f𝒦_L𝒦_R`$ and for every $`sS`$. Therefore, $$\mathrm{Col}_{sS}\stackrel{~}{U}_{[s]}^{}[M_sf]_{[s]}=\mathrm{Col}_{sS}\left[\begin{array}{cc}𝕋_{[s],L}E_{L,s}& 𝕋_{[s],R}\stackrel{~}{M}_s(T_R)\end{array}\right]f$$ (5.30) which, on account of (4.12) and (4.13) can be written as $$\mathrm{Col}_{sS}\stackrel{~}{U}_{[s]}^{}[M_sf]_{[s]}=\left[\begin{array}{cc}\left[\left(X_LH\right)^L(T_L)\right]^{}& \left(Z_{\mathrm{\Gamma },}GY_R\right)^R(T_R)\end{array}\right]f.$$ (5.31) Similarly, by (5.26), (5.27) and block partitionings (1.38) and (5.25) of $`N_r`$ and $`𝕋`$, it holds that $$[N_rf]_{[r]}=\stackrel{~}{U}_{[r]}𝕋_{[r]}(N_rf)=\stackrel{~}{U}_{[r]}\left[\begin{array}{cc}𝕋_{[r],L}\stackrel{~}{N}_r(T_L)& 𝕋_{[r],R}E_{R,r}\end{array}\right]f(rR).$$ Therefore, $$\mathrm{Col}_{rR}\stackrel{~}{U}_{[r]}^{}[N_rf]_{[r]}=\mathrm{Col}_{rR}\left[\begin{array}{cc}𝕋_{[r],L}\stackrel{~}{N}_r(T_L)& 𝕋_{[r],R}E_{R,r}\end{array}\right]f$$ (5.32) which, on account of (4.11) and (4.14) can be written as $$\mathrm{Col}_{rR}\stackrel{~}{U}_{[r]}^{}[N_rf]_{[r]}=\left[\begin{array}{cc}\left[\left(X_LHZ_{\mathrm{\Gamma },}\right)^L(T_L)\right]^{}& \left(GY_R\right)^R(T_R)\end{array}\right]f.$$ (5.33) Thus, by (5.24) and in view of (1.40), (5.31) and (5.33), $`\left[\begin{array}{cc}\stackrel{~}{A}& \stackrel{~}{B}\\ \stackrel{~}{C}& \stackrel{~}{D}\end{array}\right]\left[\begin{array}{c}\mathrm{Col}_{sS}[M_sf]_{[s]}\\ Yf\end{array}\right]`$ (5.38) $`=\left[\begin{array}{cc}_{rR}\stackrel{~}{U}_{[r]}& 0\\ 0& I\end{array}\right]\left[\begin{array}{cc}A& B\\ C& D\end{array}\right]\left[\begin{array}{c}\mathrm{Col}_{sS}\stackrel{~}{U}_{[s]}^{}[M_sf]_{[s]}\\ Yf\end{array}\right]`$ (5.45) $`=\left[\begin{array}{cc}_{rR}\stackrel{~}{U}_{[r]}& 0\\ 0& I\end{array}\right]\left[\begin{array}{cc}A& B\\ C& D\end{array}\right]\left[\begin{array}{cc}\left[\left(X_LH\right)^L(T_L)\right]^{}& \left(Z_{\mathrm{\Gamma },}GY_R\right)^R(T_R)\\ Y_L^{}& Y_R\end{array}\right]f`$ (5.50) $`=\left[\begin{array}{cc}_{rR}\stackrel{~}{U}_{[r]}& 0\\ 0& I\end{array}\right]\left[\begin{array}{cc}\left[\left(X_LHZ_{\mathrm{\Gamma },}\right)^L(T_L)\right]^{}& \left(GY_R\right)^R(T_R)\\ X_L^{}& X_R\end{array}\right]=\left[\begin{array}{c}\mathrm{Col}_{rR}[N_rf]_{[r]}\\ Xf\end{array}\right],`$ (5.53) which proves (5.29) and completes the proof of the lemma.∎ ###### Lemma 5.2. Let $`\stackrel{~}{𝐔}`$ of the form (5.8) be a unitary extension of the isometry $`𝐕`$ given in (5.4). Then the characteristic function $`F`$ of the unitary colligation of the form (5.7), $$F(z)=\stackrel{~}{D}+\stackrel{~}{C}\left(IZ_{\mathrm{\Gamma },\stackrel{~}{}}(z)\stackrel{~}{A}\right)^1Z_{\mathrm{\Gamma },\stackrel{~}{}}(z)\stackrel{~}{B},$$ is a solution to Problem 1.10. Proof: We use the arguments from the proof of the previous lemma in the reverse order. We start with positive semidefinite operators $`𝕂_p(_p^{})`$ (the spaces $`_p^{}`$ are given in (5.1)) and fix their factorizations $$𝕂_p=𝕋_p^{}𝕋_p\text{with}𝕋_p=\left[\begin{array}{cc}𝕋_{p,L}& 𝕋_{p,R}\end{array}\right]:_p^{}_p$$ (5.54) where $`=\{_p:pP\}`$ is a collection of auxiliary Hilbert spaces. Comparing (5.54) with (1.41) we get factorizations $$𝕂_{p,L}=𝕋_{p,L}^{}𝕋_{p,L},𝕂_{p,R}=𝕋_{p,R}^{}𝕋_{p,R},𝕂_{p,LR}=𝕋_{p,L}^{}𝕋_{p,R}.$$ for the block entries in $`𝕂_p`$ and more detailed decompositions (1.42) lead us to equalities $`E_{L,s}^{}𝕋_{[s],L}^{}𝕋_{[s^{}],L}E_{L,s^{}}`$ $`=`$ $`E_{L,s}^{}𝕂_{[s],L}E_{L,s^{}}=\left[𝕂_{[s],L}\right]_{s,s^{}}=\mathrm{\Psi }_{s,s^{}},`$ (5.55) $`E_{L,s}^{}𝕋_{[s],L}^{}𝕋_{[r],R}E_{L,r}`$ $`=`$ $`E_{L,s}^{}𝕂_{[s],LR}E_{R,r}=\left[𝕂_{[s],LR}\right]_{s,r}=\mathrm{\Lambda }_{s,r},`$ (5.56) $`E_{R,r}^{}𝕋_{[r],R}^{}𝕋_{[r^{}],R}E_{R,r^{}}`$ $`=`$ $`E_{R,r}^{}𝕂_{[r],R}E_{R,r^{}}=\left[𝕂_{[r],L}\right]_{r,r^{}}=\mathrm{\Phi }_{r,r^{}}`$ (5.57) (where $`E_{L,s}`$ and $`E_{R,r}`$ are given by (1.33), (1.34)) holding for every choice of $`s,s^{}S`$ and $`r,r^{}R`$ so that $`[s]=[s^{}]=[r]=[r^{}]`$. The latter equalities suggest the introduction of the operators $`𝔽_L`$ $`=`$ $`\mathrm{Col}_{sS}𝕋_{[s],L}E_{L,s}:𝒦_L{\displaystyle \underset{sS}{}}_{[s]},`$ (5.58) $`𝔽_R`$ $`=`$ $`\mathrm{Col}_{rR}𝕋_{[r],R}E_{R,r}:𝒦_R{\displaystyle \underset{rR}{}}_{[r]}.`$ (5.59) We note the following two formulas $`\mathrm{Col}_{rR}𝕋_{[r],L}\stackrel{~}{N}_r(T_L)`$ $`=`$ $`\left[\left(𝔽_L^{}Z_{\mathrm{\Gamma },}\right)^L(T_L)\right]^{},`$ (5.60) $`\mathrm{Col}_{sS}𝕋_{[s],R}\stackrel{~}{M}_s(T_R)`$ $`=`$ $`\left(Z_{\mathrm{\Gamma },}𝔽_R\right)^R(T_R),`$ (5.61) which are similar to formulas (4.11) and (4.13) and are verified in much the same way. Let $`\stackrel{~}{U}=\{\stackrel{~}{U}_p:pP\}`$ be the collection of unitary maps indexed by the set of path-connected components $`P`$ of $`\mathrm{\Gamma }`$ and defined via formulas (5.26), (5.27). Then relations (5.30) and (5.32) hold by construction; in view of (5.58)–(5.61) these relations can be written as $`\mathrm{Col}_{sS}\stackrel{~}{U}_{[s]}^{}[M_sf]_{[s]}`$ $`=`$ $`\left[\begin{array}{cc}𝔽_L& \left(Z_{\mathrm{\Gamma },}𝔽_R\right)^R(T_R)\end{array}\right]f,`$ (5.62) $`\mathrm{Col}_{rR}\stackrel{~}{U}_{[r]}^{}[N_rf]_{[r]}`$ $`=`$ $`\left[\begin{array}{cc}\left[\left(𝔽_L^{}Z_{\mathrm{\Gamma },}\right)^L(T_L)\right]^{}& 𝔽_R\end{array}\right]f.`$ (5.63) Now we define the operator $$𝐔=\left[\begin{array}{cc}A& B\\ C& D\end{array}\right]=\left[\begin{array}{cc}[A_{r,s}]& [B_r]\\ [C_s]& D\end{array}\right]:\left[\begin{array}{c}_{sS}_{[s]}\\ 𝒰\end{array}\right]\left[\begin{array}{c}_{rR}_{[r]}\\ 𝒴\end{array}\right]$$ in accordance to (5.28) by $$𝐔=\left[\begin{array}{cc}A& B\\ C& D\end{array}\right]=\left[\begin{array}{cc}_{rR}\stackrel{~}{U}_{[r]}^{}& 0\\ 0& I\end{array}\right]\left[\begin{array}{cc}\stackrel{~}{A}& \stackrel{~}{B}\\ \stackrel{~}{C}& \stackrel{~}{D}\end{array}\right]\left[\begin{array}{cc}_{sS}\stackrel{~}{U}_{[s]}& 0\\ 0& I\end{array}\right].$$ By the assumption of the lemma, $`\stackrel{~}{𝐔}`$ extends $`𝐕`$: $$\left[\begin{array}{cc}\stackrel{~}{A}& \stackrel{~}{B}\\ \stackrel{~}{C}& \stackrel{~}{D}\end{array}\right]\left[\begin{array}{c}\mathrm{Col}_{sS}[M_sf]_{[s]}\\ Yf\end{array}\right]=\left[\begin{array}{c}\mathrm{Col}_{rR}[N_rf]_{[r]}\\ Xf\end{array}\right]\text{for every}f𝒦_L𝒦_R,$$ which can be written in terms of $`𝐔`$ as $$\left[\begin{array}{cc}A& B\\ C& D\end{array}\right]\left[\begin{array}{c}\mathrm{Col}_{sS}\stackrel{~}{U}_{[s]}^{}[M_sf]_{[s]}\\ Yf\end{array}\right]=\left[\begin{array}{c}\mathrm{Col}_{rR}\stackrel{~}{U}_{[r]}^{}[N_rf]_{[r]}\\ Xf\end{array}\right](f𝒦_L𝒦_R).$$ Upon substituting equalities (5.62) and (5.63) and block decompositions (1.40) for $`X`$ and $`Y`$ in the latter equality we get $$\left[\begin{array}{cc}A& B\\ C& D\end{array}\right]\left[\begin{array}{cc}𝔽_L& \left(Z_{\mathrm{\Gamma },}𝔽_R\right)^R(T_R)\\ Y_L^{}& Y_R\end{array}\right]=\left[\begin{array}{cc}\left[\left(𝔽_L^{}Z_{\mathrm{\Gamma },}\right)^L(T_L)\right]^{}& 𝔽_R\\ X_L^{}& X_R\end{array}\right].$$ (5.64) By Remark 2.1, the colligations $`𝚺`$ and $`\stackrel{~}{𝚺}`$ defined in (2.1) and (5.7) have the same characteristic functions and thus, $`F`$ can be taken in the form (1.20). Let $`H(z)`$ and $`G(z)`$ be defined as in (1.26) and decomposed as in (1.21) and (1.23). We shall use the representations (5.9) of $`F(z)`$ which are equivalent to (1.20). Since $`𝐔`$ is unitary, it follows from (5.64) that $`A^{}\left[\left(𝔽_L^{}Z_{\mathrm{\Gamma },}\right)^L(T_L)\right]^{}+C^{}X_L^{}`$ $`=`$ $`𝔽_L,`$ (5.65) $`B^{}\left[\left(𝔽_L^{}Z_{\mathrm{\Gamma },}\right)^L(T_L)\right]^{}+D^{}X_L^{}`$ $`=`$ $`Y_L^{},`$ (5.66) $`A\left(Z_{\mathrm{\Gamma },}𝔽_R\right)^R(T_R)+BY_R`$ $`=`$ $`𝔽_R,`$ (5.67) $`C\left(Z_{\mathrm{\Gamma },}𝔽_R\right)^R(T_R)+DY_R`$ $`=`$ $`X_R.`$ (5.68) Taking adjoints in (5.65) we get $$X_LC=𝔽_L^{}\left(𝔽_L^{}Z_{\mathrm{\Gamma },}\right)^L(T_L)A$$ which can be written, by properties (4.1) and (4.2) of the left evaluation map, as $$X_LC=\left(𝔽_L^{}(IZ_{\mathrm{\Gamma },}A)\right)^L(T_L).$$ Multiplying both sides in the last equality by $`(IZ_{\mathrm{\Gamma },}(z)A)^1`$ on the right and applying the left evaluation map to the resulting identity $$X_LH(z)=\left(𝔽_L^{}(IZ_{\mathrm{\Gamma },}A)\right)^L(T_L)(IZ_{\mathrm{\Gamma },}(z)A)^1,$$ we get $`\left(X_LH\right)^L(T_L)`$ $`=`$ $`\left(\left(𝔽_L^{}(IZ_{\mathrm{\Gamma },}A)\right)^L(T_L)(IZ_{\mathrm{\Gamma },}A)^1\right)^L(T_L)`$ (5.69) $`=`$ $`\left(𝔽_L^{}(IZ_{\mathrm{\Gamma },}A)(IZ_{\mathrm{\Gamma },}A)^1\right)^L(T_L)`$ $`=`$ $`\left(𝔽_L^{}\right)^L(T_L)=𝔽_L^{}.`$ Note that the second equality in the last chain has been obtained upon applying (4.4) to $$T(z)=𝔽_L^{}(IZ_{\mathrm{\Gamma },}(z)A)\text{and}\stackrel{~}{T}(z)=(IZ_{\mathrm{\Gamma },}(z)A)^1,$$ whereas the third equality follows by the property (4.1). Next we take adjoints in (5.66) to get $$Y_L=\left(𝔽_L^{}Z_{\mathrm{\Gamma },}\right)^L(T_L)B+X_LD=\left(𝔽_L^{}Z_{\mathrm{\Gamma },}B\right)^L(T_L)+X_LD.$$ (5.70) By (5.69), $$\left(𝔽_L^{}Z_{\mathrm{\Gamma },}B\right)^L(T_L)=\left(\left(X_LH\right)^L(T_L)Z_{\mathrm{\Gamma },}B\right)^L(T_L)$$ and applying (4.4) to $$T(z)=X_LH(z)\text{and}\stackrel{~}{T}(z)=Z_{\mathrm{\Gamma },}(z)B$$ leads us to $$\left(𝔽_L^{}Z_{\mathrm{\Gamma },}B\right)^L(T_L)=((X_LHZ_{\mathrm{\Gamma },}B)^L(T_L).$$ Substituting the latter equality into the left hand side expression in (5.70) and making use of the first representation of $`S`$ in (5.9), we get $`Y_L`$ $`=`$ $`\left(X_LHZ_{\mathrm{\Gamma },}B\right)^L(T_L)+X_LD`$ $`=`$ $`\left(X_LHZ_{\mathrm{\Gamma },}B+X_LD\right)^L(T_L)=\left(X_LS\right)^L(T_L),`$ which proves the first interpolation condition in (1.32). To get the second interpolation condition in (1.32) write (5.67) in the form $$BY_R=(IAZ_{\mathrm{\Gamma },})𝔽_R)^R(T_R),$$ multiply the latter equality by $`(IAZ_{\mathrm{\Gamma },}(z))^1`$ on the left and apply the right evaluation map to the resulting identity $$G(z)Y_R=(IAZ_{\mathrm{\Gamma },}(z))^1(IAZ_{\mathrm{\Gamma },})𝔽_R)^R(T_R).$$ We have $`\left(GY_R\right)^R(T_R)`$ $`=`$ $`\left((IAZ_{\mathrm{\Gamma },})^1\left((IAZ_{\mathrm{\Gamma },})𝔽_R\right)^R(T_R)\right)^R(T_R)`$ (5.71) $`=`$ $`\left((IAZ_{\mathrm{\Gamma },})^1(IAZ_{\mathrm{\Gamma },})𝔽_R\right)^R(T_R)`$ $`=`$ $`\left(𝔽_R\right)^R(T_R)=𝔽_R.`$ Note that the third equality in the last chain has been obtained upon applying (4.5) to $$T(z)=(IAZ_{\mathrm{\Gamma },}(z))^1\text{and}\stackrel{~}{T}(z)=(IAZ_{\mathrm{\Gamma },}(z))𝔽_R.$$ Substituting (5.71) into (5.68) and applying (4.5) to $$T(z)=CZ_{\mathrm{\Gamma },}(z)\text{and}\stackrel{~}{T}(z)=G(z)Y_R,$$ we get $`X_R`$ $`=`$ $`\left(CZ_{\mathrm{\Gamma },}\left(GY_R\right)^R(T_R)\right)^R(T_R)+DY_R`$ $`=`$ $`\left(CZ_{\mathrm{\Gamma },}GY_R\right)^R(T_R)+DY_R`$ $`=`$ $`\left(CZ_{\mathrm{\Gamma },}GY_R+DY_R\right)^R(T_R)`$ which coincides with the second equality in (1.32), due to the second representation in (5.9). Thus, $`F`$ belongs to $`𝒮𝒜_\mathrm{\Gamma }(𝒰,𝒴)`$ as the characteristic function of a unitary colligation (2.1) and satisfies interpolation conditions (1.32). It remains to show that it satisfies also conditions (1.44)–(1.46). But it follows from (5.69), (5.71) and (5.58) that $$\left(X_LH_s\right)^L(T_L)=E_{L,s}^{}𝕋_{[s],L}^{}\text{and}\left(G_rY_R\right)^L(T_R)=𝕋_{[r],R}E_{R,r}$$ for $`sS`$ and $`rR`$. Now we pick any $`s,s^{}S`$ and $`r,r^{}R`$ so that $`[s]=[s^{}]=[r]=[r^{}]`$ and combine the two latter equalities with (5.55)–(5.57) to get (1.44)–(1.46): $`(X_LH_s)^L(T_L)\left[(X_LH_s^{})^L(T_L)\right]^{}`$ $`=`$ $`E_{L,s}^{}𝕋_{[s],L}^{}𝕋_{[s^{}],L}E_{L,s^{}}=\mathrm{\Psi }_{s,s^{}},`$ $`(X_LH_s)^L(T_L)\left(G_rY_R\right)^R(T_R)`$ $`=`$ $`E_{L,s}^{}𝕋_{[s],L}^{}𝕋_{[r],R}E_{R,r}=\mathrm{\Lambda }_{s,r},`$ $`\left[(G_rY_R)^R(T_R)\right]^{}(G_r^{}Y_R)^R(T_R)`$ $`=`$ $`E_{R,r}^{}𝕋_{[r],R}^{}𝕋_{[r^{}],R}E_{R,r^{}}=\mathrm{\Phi }_{r,r^{}},`$ and complete the proof.∎ ## 6. The universal unitary colligation associated with the interpolation problem A general result of Arov and Grossman (see , ) describes how to parametrize the set of all unitary extensions of a given partially defined isometry $`𝐕`$. Their result has been extended to the multivariable (commutative) case in and will be extended in this section to the setting of noncommutative power series. Let $`𝐕:𝒟_𝐕_𝐕`$ be the isometry given in (5.4) with $`𝒟_𝐕`$ and $`_𝐕`$ given in (5.5) and (5.6). Introduce the defect spaces $$\mathrm{\Delta }=\left[\begin{array}{c}_{sS}\widehat{}_{[s]}\\ 𝒰\end{array}\right]𝒟_𝐕\mathrm{and}\mathrm{\Delta }_{}=\left[\begin{array}{c}_{rR}\widehat{}_{[r]}\\ 𝒴\end{array}\right]_𝐕$$ and let $`\stackrel{~}{\mathrm{\Delta }}`$ to be another copy of $`\mathrm{\Delta }`$ and $`\stackrel{~}{\mathrm{\Delta }}_{}`$ to be another copy of $`\mathrm{\Delta }_{}`$ with unitary identification maps $$i:\mathrm{\Delta }\stackrel{~}{\mathrm{\Delta }}\text{and}i_{}:\mathrm{\Delta }_{}\stackrel{~}{\mathrm{\Delta }}_{}.$$ Define a unitary operator $`𝐔_0`$ from $`𝒟_𝐕\mathrm{\Delta }\stackrel{~}{\mathrm{\Delta }}_{}`$ onto $`_𝐕\mathrm{\Delta }_{}\stackrel{~}{\mathrm{\Delta }}`$ by the rule $$𝐔_0x=\{\begin{array}{cc}𝐕x,\hfill & \text{if}x𝒟_𝐕,\hfill \\ i(x)\hfill & \text{if}x\mathrm{\Delta },\hfill \\ i_{}^1(x)\hfill & \text{if}x\stackrel{~}{\mathrm{\Delta }}_{}.\hfill \end{array}$$ (6.1) Identifying $`\left[\begin{array}{c}𝒟_𝐕\\ \mathrm{\Delta }\end{array}\right]`$ with $`\left[\begin{array}{c}_{sS}\widehat{}_{[s]}\\ 𝒰\end{array}\right]`$ and $`\left[\begin{array}{c}_𝐕\\ \mathrm{\Delta }_{}\end{array}\right]`$ with $`\left[\begin{array}{c}_{rR}\widehat{}_{[r]}\\ 𝒴\end{array}\right]`$, we decompose $`𝐔_0`$ defined by (6.1) according to $$𝐔_0=\left[\begin{array}{ccc}U_{11}& U_{12}& U_{13}\\ U_{21}& U_{22}& U_{23}\\ U_{31}& U_{32}& 0\end{array}\right]:\left[\begin{array}{c}_{sS}\widehat{}_{[s]}\\ 𝒰\\ \stackrel{~}{\mathrm{\Delta }}_{}\end{array}\right]\left[\begin{array}{c}_{rR}\widehat{}_{[r]}\\ 𝒴\\ \stackrel{~}{\mathrm{\Delta }}\end{array}\right].$$ (6.2) The $`(3,3)`$ block in this decomposition is zero, since (by definition (6.1)), for every $`x\stackrel{~}{\mathrm{\Delta }}_{}`$, the vector $`𝐔_0x`$ belongs to $`\mathrm{\Delta }`$, which is a subspace of $`\left[\begin{array}{c}_{rR}\widehat{}_{[r]}\\ 𝒴\end{array}\right]`$ and therefore, is orthogonal to $`\stackrel{~}{\mathrm{\Delta }}`$ (in other words $`𝐏_{\stackrel{~}{\mathrm{\Delta }}}𝐔_0|_{\stackrel{~}{\mathrm{\Delta }}_{}}=0`$, where $`𝐏_{\stackrel{~}{\mathrm{\Delta }}}`$ stands for the orthogonal projection of $`_𝐕\mathrm{\Delta }_{}\stackrel{~}{\mathrm{\Delta }}`$ onto $`\stackrel{~}{\mathrm{\Delta }}`$). The unitary operator $`𝐔_0`$ is the connecting operator of the unitary colligation $$𝚺_0=\{\mathrm{\Gamma },\widehat{},\left[\begin{array}{c}𝒰\\ \stackrel{~}{\mathrm{\Delta }}_{}\end{array}\right],\left[\begin{array}{c}𝒴\\ \stackrel{~}{\mathrm{\Delta }}\end{array}\right],𝐔_0\},$$ (6.3) which is called the universal unitary colligation associated with the interpolation Problem 1.10. Let $`\stackrel{~}{𝚺}`$ be any colligation of the form $$\stackrel{~}{𝚺}=\{\mathrm{\Gamma },\stackrel{~}{},\stackrel{~}{\mathrm{\Delta }},\stackrel{~}{\mathrm{\Delta }}_{},\stackrel{~}{𝐔}\}.$$ (6.4) We define another colligation $`_{𝚺_0}[\stackrel{~}{𝚺}]`$, called the coupling of $`𝚺_0`$ and $`\stackrel{~}{𝚺}`$, to be the colligation of the form $$_{𝚺_0}[\stackrel{~}{𝚺}]=\{\mathrm{\Gamma },\widehat{}\stackrel{~}{},𝒰,𝒴,_{𝐔_0}[\stackrel{~}{𝐔}]\}$$ with the connecting operator $`_{𝐔_0}[\stackrel{~}{𝐔}]`$ defined as follows: $$_{𝐔_0}[\stackrel{~}{𝐔}]:\left[\begin{array}{c}c\\ h\\ u\end{array}\right]\left[\begin{array}{c}c^{}\\ h^{}\\ y\end{array}\right]$$ (6.5) if the system of equations $$𝐔_0:\left[\begin{array}{c}c\\ u\\ \stackrel{~}{d}_{}\end{array}\right]\left[\begin{array}{c}c^{}\\ y\\ \stackrel{~}{d}\end{array}\right]\text{and}\stackrel{~}{𝐔}:\left[\begin{array}{c}h\\ \stackrel{~}{d}\end{array}\right]\left[\begin{array}{c}h^{}\\ \stackrel{~}{d}_{}\end{array}\right]$$ (6.6) is satisfied for some choice of $`\stackrel{~}{d}\stackrel{~}{\mathrm{\Delta }}`$ and $`\stackrel{~}{d}_{}\stackrel{~}{\mathrm{\Delta }}_{}`$. To show that the operator $`_{𝐔_0}[\stackrel{~}{𝐔}]`$ is well defined, i.e., that for every triple $`(c,h,u)`$, there exist $`\stackrel{~}{d}`$ and $`\stackrel{~}{d}_{}`$ for which the system (6.6) is consistent and the resulting triple $`(c^{},h^{},y)`$ does not depend on the choice of $`\stackrel{~}{d}`$ and $`\stackrel{~}{d}_{}`$, we note first that, on account of (6.1) and (6.2), the the bottom component of the first equation in (6.6) determines $`\stackrel{~}{d}`$ uniquely by $$\stackrel{~}{d}=𝐏_{\stackrel{~}{\mathrm{\Delta }}}\left(\mathrm{𝐕𝐏}_{𝒟_𝐕}+i𝐏_\mathrm{\Delta }\right)\left[\begin{array}{c}c\\ u\end{array}\right]=i𝐏_\mathrm{\Delta }\left[\begin{array}{c}c\\ u\end{array}\right].$$ With this $`\stackrel{~}{d}`$, the the bottom component of the second equation in (6.6) determines uniquely $`\stackrel{~}{d}_{}`$ and $`h^{}`$. Using $`\stackrel{~}{d}_{}`$ one can recover now $`c^{}`$ and $`y`$ from the first and second components of the first equation in (6.6). Since operators $`𝐔_0`$ and $`\stackrel{~}{𝐔}`$ are unitary, it follows from (6.6) that $`c^2+u^2+\stackrel{~}{d}_{}^2`$ $`=`$ $`c^{}^2+y^2+\stackrel{~}{d}^2,`$ $`h^2+\stackrel{~}{d}^2`$ $`=`$ $`h^{}^2+\stackrel{~}{d}_{}^2`$ and therefore, that $$c^2+u^2+h^2=c^{}^2+y^2+h^{}^2,$$ which means that the coupling operator $`_{𝐔_0}[\stackrel{~}{𝐔}]`$ is isometric. A similar argument can be made with the adjoints of $`𝐔_0`$, $`\stackrel{~}{𝐔}`$ and $`_{𝐔_0}[\stackrel{~}{𝐔}]`$, and hence $`_{𝐔_0}[\stackrel{~}{𝐔}]`$ is unitary. Furthermore, by (6.5) and (6.6), $$_{𝐔_0}[\stackrel{~}{𝐔}]|_{(_{sS}\widehat{}_{[s]})𝒰}=𝐔_0|_{(_{sS}\widehat{}_{[s]})𝒰}$$ and since $`𝒟_𝐕(_{sS}\widehat{}_{[s]})𝒰`$, it follows that $$_{𝐔_0}[\stackrel{~}{𝐔}]|_{𝒟_𝐕}=𝐔_0|_{𝒟_𝐕}=𝐕.$$ (6.7) Thus, the coupling of the connecting operator $`𝐔_0`$ of the universal unitary colligation associated with Problem 1.10 and any other unitary operator is a unitary extension of the isometry $`𝐕`$ defined in (5.4). Conversely for every unitary colligation $`𝚺=\{\mathrm{\Gamma },\widehat{}\stackrel{~}{},𝒰,𝒴,𝐔\}`$ with the connecting operator being a unitary extension of $`𝐕`$, there exists a unitary colligation $`\stackrel{~}{𝚺}`$ of the form (6.4) such that $`𝚺=_{𝚺_0}[\stackrel{~}{𝚺}]`$ (the proof is the same as in \[22, Theorem 6.2\]). Thus, all unitary extensions $`𝐔`$ of the isometry $`𝐕`$ defined in (5.4) are parametrized by the formula $$𝐔=_{𝐔_0}[\stackrel{~}{𝐔}],\stackrel{~}{𝐔}:(_{sS}\widehat{}_{[s]})\stackrel{~}{\mathrm{\Delta }}(_{rR}\widehat{}_{[r]})\stackrel{~}{\mathrm{\Delta }}_{}$$ (6.8) and $`\stackrel{~}{}=\{\stackrel{~}{}:pP\}`$ is a collection of auxiliary Hilbert spaces indexed by the path-connected components $`pP=P(\mathrm{\Gamma })`$ of the admissible graph $`\mathrm{\Gamma }`$. According to (2), the characteristic function of the colligation $`𝚺_0`$ defined in (6.3) with the connecting operator $`𝐔_0`$ partitioned as in (6.2), is given by $`\mathrm{\Sigma }(z)`$ $`=`$ $`\left[\begin{array}{cc}\mathrm{\Sigma }_{11}(z)& \mathrm{\Sigma }_{12}(z)\\ \mathrm{\Sigma }_{21}(z)& \mathrm{\Sigma }_{22}(z)\end{array}\right]`$ (6.11) $`=`$ $`\left[\begin{array}{cc}U_{22}& U_{23}\\ U_{32}& 0\end{array}\right]+\left[\begin{array}{c}U_{21}\\ U_{31}\end{array}\right]\left(IZ_{\mathrm{\Gamma },\widehat{}}(z)U_{11}\right)^1Z_{\mathrm{\Gamma },\widehat{}}(z)\left[\begin{array}{cc}U_{12}& U_{13}\end{array}\right]`$ (6.16) and belongs to the class $`𝒮𝒜_\mathrm{\Gamma }(𝒰\stackrel{~}{\mathrm{\Delta }}_{},𝒴\stackrel{~}{\mathrm{\Delta }})`$ by Theorem 1.7. ###### Theorem 6.1. Let $`𝐕`$ be the isometry defined in (5.4), let $`\mathrm{\Sigma }`$ be constructed as above and let $`F`$ be an element in $`(𝒰,𝒴)z`$. Then the following are equivalent: 1. $`F`$ is a solution of Problem 1.10. 2. $`F`$ is a characteristic function of a colligation $`𝚺=\{\mathrm{\Gamma },\widehat{}\stackrel{~}{},𝒰,𝒴,𝐔\}`$ with the connecting operator $`𝐔`$ being a unitary extension of $`𝐕`$. 3. $`F`$ is of the form $$F(z)=\mathrm{\Sigma }_{11}(z)+\mathrm{\Sigma }_{12}(z)\left(I_{\stackrel{~}{\mathrm{\Delta }}_{}}𝒯(z)\mathrm{\Sigma }_{22}(z)\right)^1𝒯(z)\mathrm{\Sigma }_{21}(z)$$ (6.17) where $`𝒯`$ is a power series from the noncommutative Schur-Agler class $`𝒮𝒜_\mathrm{\Gamma }(\stackrel{~}{\mathrm{\Delta }},\stackrel{~}{\mathrm{\Delta }}_{})`$. Proof: The equivalence $`\mathrm{𝟏}\mathrm{𝟐}`$ follows by Lemmas 5.1 and 5.2. $`\mathrm{𝟐}\mathrm{𝟑}`$. By the preceding analysis, the colligation $`𝚺`$ is the coupling of the universal colligation $`𝚺_0`$ defined in (6.3) and some unitary colligation $`\stackrel{~}{𝚺}`$ of the form (6.4). The connecting operators $`𝐔`$, $`𝐔_0`$ and $`\stackrel{~}{𝐔}`$ of these colligations are related as in (6.8). Let $`F`$, $`\mathrm{\Sigma }`$ and $`𝒯`$ be characteristic functions of $`𝚺`$, $`𝚺_0`$ and $`\stackrel{~}{𝚺}`$, respectively. Applying Remark 2.2 to (6.5) and (6.6), we get $$F(z)e=e_{},\mathrm{\Sigma }(z)\left[\begin{array}{c}u\\ \stackrel{~}{d}_{}\end{array}\right]=\left[\begin{array}{c}y\\ \stackrel{~}{d}\end{array}\right],𝒯(z)\stackrel{~}{d}=\stackrel{~}{d}_{}.$$ (6.18) Substituting the third relation in (6.18) into the second we get $$\mathrm{\Sigma }(z)\left[\begin{array}{c}u\\ 𝒯(z)\stackrel{~}{d}\end{array}\right]=\left[\begin{array}{c}y\\ \stackrel{~}{d}\end{array}\right],$$ which in view of the block decomposition (6.16) of $`\mathrm{\Sigma }`$ splits into $$\mathrm{\Sigma }_{11}(z)u+\mathrm{\Sigma }_{12}(z)𝒯(z)\stackrel{~}{d}=y\text{and}\mathrm{\Sigma }_{21}(z)u+\mathrm{\Sigma }_{22}(z)𝒯(z)\stackrel{~}{d}=\stackrel{~}{d}.$$ The second from the two last equalities gives $$\stackrel{~}{d}=\left(I\mathrm{\Sigma }_{22}(z)𝒯(z)\right)^1\mathrm{\Sigma }_{21}(z)u$$ which, being substituted into the first equality, implies $$\left(\mathrm{\Sigma }_{11}(z)+\mathrm{\Sigma }_{12}(z)𝒯(z)\left(I\mathrm{\Sigma }_{22}(z)𝒯(z)\right)^1\mathrm{\Sigma }_{21}(z)\right)u=y.$$ The latter is equivalent to $$\left(\mathrm{\Sigma }_{11}(z)+\mathrm{\Sigma }_{12}(z)\left(I𝒯(z)\mathrm{\Sigma }_{22}(z)\right)^1𝒯(z)\mathrm{\Sigma }_{21}(z)\right)u=y$$ and the comparison of the last equality with the first relation in (6.18) leads to representation (6.17) of $`F`$, since a vector $`u𝒰`$ is arbitrary. $`\mathrm{𝟑}\mathrm{𝟐}`$. Let $`F`$ be of the form (6.17) for some $`𝒯𝒮𝒜_\mathrm{\Gamma }(\stackrel{~}{\mathrm{\Delta }},\stackrel{~}{\mathrm{\Delta }}_{})`$. By Theorem 1.7, $`𝒯`$ is the characteristic function of a unitary colligation $`\stackrel{~}{𝚺}`$ of the form (6.4). Let $`𝚺`$ be the unitary colligation defined by $`𝚺=_{𝚺_0}[\stackrel{~}{𝚺}]`$. By the preceding “$`\mathrm{𝟐}\mathrm{𝟑}`$” part, $`F`$ of the form (6.17) is the characteristic function of $`𝚺`$. It remains to note that the colligation $`𝚺`$ is of required the form: its input and output spaces coincide with $`𝒰`$ and $`𝒴`$, respectively (by the definition of coupling) and its connecting operator is an extension of $`𝐕`$, by (6.7).∎ As a corollary we obtain the sufficiency part in both Theorem 1.40 and Theorem 1.11, including the parametrization of the set of all solutions of Problem 1.10 in Theorem 1.11 and the parametrization of the set of all solutions of Problem 1.8 in Corollary 1.12. ## 7. Examples and special cases For certain special cases of Problems 1.8 and 1.10, the general interpolation results stated in Theorems 1.40 and 1.11 become much more transparent. Moreover, some of these particular cases are quite important for applications and are interesting in their own right; it seems reasonable therefore to display them in more detail. ### 7.1. Left sided interpolation problems The left sided problem can be considered as the special case of Problem 1.8 when $`T_R`$ is a tuple of operators acting on the space of dimension zero. ###### Problem 7.1. Given an admissible data set $`𝒟=\{T_L,X_L,Y_L\}`$, find necessary and sufficient conditions for existence of a power series $`F𝒮𝒜_\mathrm{\Gamma }(𝒰,𝒴)`$ such that $$\left(X_LF\right)^L(T_L)=Y_L.$$ (7.1) The answer follows immediately from Theorem 1.40. ###### Theorem 7.2. There is a power series $`F𝒮𝒜_\mathrm{\Gamma }(𝒰,𝒴)`$ satisfying interpolation condition (7.1) if and only if there exists a collection $`𝕂_L=\{𝕂_{p,L}:pP\}`$ of positive semidefinite operators on the space $`_{sS:[s]=p}𝒦_L`$ indexed by the set of path-connected components $`P`$ of $`\mathrm{\Gamma }`$, which satisfies the Stein identity $$\underset{sS}{}E_{L,s}^{}𝕂_{[s],L}E_{L,s}\underset{rR}{}\stackrel{~}{N}_r(T_L)^{}𝕂_{[r],L}\stackrel{~}{N}_r(T_L)=X_LX_L^{}Y_LY_L^{},$$ (7.2) where $`E_{L,s}`$ and $`\stackrel{~}{N}_r`$ are the operators defined via formulas (1.33) and (1.35), respectively. Furthermore, it follows by Theorem 1.11 that for every choice of a tuple $`𝕂_L`$ satisfying the conditions of Theorem 7.2, there exists a power series $`F𝒮𝒜_\mathrm{\Gamma }(𝒰,𝒴)`$ satisfying (besides the left interpolation condition (7.1)) supplementary interpolation conditions $$(X_LH_s)^L(T_L)\left[(X_LH_s^{})^L(T_L)\right]^{}=\mathrm{\Psi }_{s,s^{}}\text{ for }s,s^{}S:[s]=[s^{}],$$ (7.3) for some choice of associated function $`H(z)`$ in representation (1.22) of $`F`$. Furthermore, all such $`F`$ can be parametrized by a linear fractional transformation. We leave to the reader to formulate the right sided interpolation problem and to derive the right sided version of Theorem 7.2 from Theorem 1.40. Parallel results hold for right-sided interpolation problems; we leave the formulation of explicit statements to the reader. ### 7.2. The case of the noncommutative ball Now we consider the Fornasini-Marchesini case (see Example 1.2 above) where $`S=\{1\}`$ and $`R=E=\{1,\mathrm{},d\}`$. In this case, from Corollaries 3.6 and 3.8 we see that a sufficient condition for $`T_L=(T_{L,1},\mathrm{},T_{L,d})`$ to be left-admissible is that $`T_L`$ be a strict row contraction and that a sufficient condition for $`T_R=(T_{R,1},\mathrm{},T_{R,d})`$ to be right admissible is that $`T_R`$ be a strict column contraction: $$\underset{j=1}{\overset{d}{}}T_{L,j}T_{L,j}^{}<I_{𝒦_L}\text{and}\underset{j=1}{\overset{d}{}}T_{R,j}^{}T_{R,j}<I_{𝒦_R}.$$ The left sided problem is of special interest. ###### Problem 7.3. Given an admissible data set $`𝒟=\{T_L,X_L,Y_L\}`$, find necessary and sufficient conditions for existence of a power series $`F𝒮𝒜_{\mathrm{\Gamma }^{\mathrm{FM}}}(𝒰,𝒴)`$ satisfying the left sided interpolation condition (7.1). In this particular case $$E_L=I_{𝒦_L},\stackrel{~}{N}_j(T_L)=T_{L,j}^{}$$ (7.4) and we conclude by Theorem 7.2 that there exists a power series $`F𝒮𝒜_{\mathrm{\Gamma }^{\mathrm{FM}}}(𝒰,𝒴)`$ satisfying (7.1) if and only if there exists a positive semidefinite operator $`𝕂_L`$ subject to the Stein identity $$𝕂_L\underset{j=1}{\overset{d}{}}T_{L,j}𝕂_LT_{L,j}^{}=X_LX_L^{}Y_LY_L^{}.$$ Since the $`d`$-tuple $`T_L`$ is a strict row contraction, the latter Stein equation has a unique solution given in terms of convergent series by $$𝕂_L=\underset{v_E}{}T_L^v\left(X_LX_L^{}Y_LY_L^{}\right)(T_L^{})^v$$ (7.5) and we come to the following. ###### Theorem 7.4. Assume that $`T_L=(T_{L,1},\mathrm{},T_{L,d})`$ is a strict row contraction. Then there is a power series $`F𝒮𝒜_{\mathrm{\Gamma }^{\mathrm{FM}}}(𝒰,𝒴)`$ satisfying interpolation condition (7.1) if and only if the operator $`𝕂_L`$ defined in (7.5) is positive semidefinite. A remarkable part about the left-sided interpolation for the Fornasini-Marchesini case is that no supplementary conditions are needed to get a parametrization of the solution set: since the operator $`𝕂_L`$ is uniquely determined by the interpolation date, it follows by Theorem 1.11 that for every $`F𝒮𝒜_{\mathrm{\Gamma }^{\mathrm{FM}}}(𝒰,𝒴)`$ satisfying (7.1), the function $`H(z)`$ associated with $`F`$ via representation (1.22), satisfies $$(X_LH)^L(T_L)\left[(X_LH)^L(T_L)\right]^{}=𝕂_L.$$ Furthermore, in this case the power series $`\mathrm{\Sigma }`$ defined in (1.48) depends on the data $`\{T_L,X_L,Y_L\}`$ only and the linear fractional formula (1.49) parametrizes the solution set to Problem 7.3. The two sided problem in the Fornasini-Marchesini case is less remarkable. ###### Problem 7.5. Given an admissible interpolation data set (1.31), find necessary and sufficient conditions for existence of a power series $`F𝒮𝒜_{\mathrm{\Gamma }^{\mathrm{FM}}}(𝒰,𝒴)`$ such that $$\left(X_LF\right)^L(T_L)=Y_L\text{and}\left(FY_R\right)^R(T_R)=X_R.$$ (7.6) The formulas (1.37) and (1.38) read $$M=\left[\begin{array}{cc}I_{𝒦_L}& 0\\ 0& T_{R,1}\\ \mathrm{}& \mathrm{}\\ 0& T_{R,d}\end{array}\right]\text{and}N_j=\left[\begin{array}{cc}T_{L,j}^{}& 0\\ 0& E_j\end{array}\right](j=1,\mathrm{},d)$$ (7.7) where $$E_1=\left[\begin{array}{c}I_{𝒦_R}\\ 0\\ \mathrm{}\\ 0\end{array}\right],E_2=\left[\begin{array}{c}0\\ I_{𝒦_R}\\ \mathrm{}\\ 0\end{array}\right],\mathrm{},E_d=\left[\begin{array}{c}0\\ \mathrm{}\\ 0\\ I_{𝒦_R}\end{array}\right].$$ Now Theorem 1.40 leads us to the following conclusion: ###### Theorem 7.6. There is a power series $`F𝒮𝒜_{\mathrm{\Gamma }^{\mathrm{FM}}}(𝒰,𝒴)`$ satisfying interpolation conditions (7.6) if and only if there exists a positive semidefinite operator $$𝕂=\left[\begin{array}{cc}𝕂_L& 𝕂_{LR}\\ 𝕂_{LR}^{}& 𝕂_R\end{array}\right](𝒦_L𝒦_R^d)$$ (7.8) subject to the Stein identity $$M^{}𝕂M\underset{j=1}{\overset{d}{}}N_j^{}𝕂N_j=X^{}XY^{}Y,$$ (7.9) where $`M`$, $`N_j`$, $`X`$ and $`Y`$ are defined in (7.7) and (1.40). Since the block $`𝕂_L`$ in (7.8) is uniquely determined from the left interpolation data via the Stein identity (7.9), the latter result can be displayed more explicitly in terms of a structured positive completion problem. ###### Theorem 7.7. There is a power series $`F𝒮𝒜_{\mathrm{\Gamma }^{\mathrm{FM}}}(𝒰,𝒴)`$ satisfying interpolation conditions (7.6) if and only if there exist operators $`\mathrm{\Lambda }_j(𝒦_R,𝒦_L)`$ and $`\mathrm{\Phi }_{ij}(𝒦_R)`$ for $`i,j=1,\mathrm{},d`$ subject to Stein identities $`{\displaystyle \underset{j=1}{\overset{d}{}}}(T_{L,j}\mathrm{\Lambda }_j\mathrm{\Lambda }_jT_{R,j})`$ $`=`$ $`Y_LY_RX_LX_R,`$ (7.10) $`{\displaystyle \underset{j=1}{\overset{d}{}}}\mathrm{\Phi }_{jj}{\displaystyle \underset{i,j=1}{\overset{d}{}}}T_{R,i}^{}\mathrm{\Phi }_{ij}T_{R,j}`$ $`=`$ $`Y_R^{}Y_RX_R^{}X_R`$ (7.11) and such that the operator $$\left[\begin{array}{cccc}𝕂_L& \mathrm{\Lambda }_1& \mathrm{}& \mathrm{\Lambda }_d\\ \mathrm{\Lambda }_1^{}& \mathrm{\Phi }_{11}& \mathrm{}& \mathrm{\Phi }_{1d}\\ \mathrm{}& \mathrm{}& & \mathrm{}\\ \mathrm{\Lambda }_d^{}& \mathrm{\Phi }_{d1}& \mathrm{}& \mathrm{\Phi }_{dd}\end{array}\right]$$ is positive semidefinite, where $`𝕂_L`$ is defined in (7.5). To get Theorem 7.7 from Theorem 7.6, it suffices to let $`𝕂_{LR}=\left[\begin{array}{ccc}\mathrm{\Lambda }_1& \mathrm{}& \mathrm{\Lambda }_d\end{array}\right]`$ and $`𝕂_R=[\mathrm{\Phi }_{ij}]_{i,j=1}^d`$ and to make use of block decompositions (7.7) and (7.8). ### 7.3. The case of the noncommutative polydisk Here we consider the Givone-Roesser case (see Example 1.3 above) where $`S=R=E=\{1,\mathrm{},d\}`$ and the tuples $`T_L`$ and $`T_R`$ are just $`d`$-tuples $`T_L=(T_{L,1},\mathrm{},T_{L,d})`$ and $`T_R=(T_{R,1},\mathrm{},T_{R,d})`$ of contractive operators acting on $`𝒦_L`$ and $`𝒦_R`$, respectively. ###### Problem 7.8. Given an admissible interpolation data set (1.31), find necessary and sufficient conditions for existence of a power series $`F𝒮𝒜_{\mathrm{\Gamma }^{\mathrm{GR}}}(𝒰,𝒴)`$ such that $$\left(X_LF\right)^L(T_L)=Y_L\text{and}\left(FY_R\right)^R(T_R)=X_R.$$ (7.12) The formulas (1.33)–(1.36) read $$E_{L,j}=I_{𝒦_L},E_{R,j}=I_{𝒦_R},\stackrel{~}{N}_j(T_L)=T_{L,j}^{},\stackrel{~}{M}_j(T_R)=T_{R,j}(j=1,\mathrm{},d)$$ and therefore, formulas (1.37) and (1.38) take the form $$M_j=\left[\begin{array}{cc}I_{𝒦_L}& 0\\ 0& T_{R,j}\end{array}\right],N_j=\left[\begin{array}{cc}T_{L,j}^{}& 0\\ 0& I_{𝒦_R}\end{array}\right](j=1,\mathrm{},d).$$ (7.13) Theorem 1.40 now reduces to ###### Theorem 7.9. There is a power series $`F𝒮𝒜_\mathrm{\Gamma }^{\mathrm{GR}}(𝒰,𝒴)`$ satisfying interpolation conditions (7.12) if and only if there exist positive semidefinite operators $$𝕂_j=\left[\begin{array}{cc}𝕂_{j,L}& 𝕂_{j,LR}\\ 𝕂_{j,LR}^{}& 𝕂_{j,R}\end{array}\right](𝒦_L𝒦_R)\text{for}j=1,\mathrm{},d,$$ (7.14) that satisfy the Stein identity $$\underset{j=1}{\overset{d}{}}\left(M_j^{}𝕂_jM_jN_j^{}𝕂_jN_j\right)=X^{}XY^{}Y,$$ (7.15) where $`M_j`$ and $`N_j`$ are the operators defined via formulas (7.13) and $`X`$ and $`Y`$ are the same as in (1.40). Furthermore, it follows by Theorem 1.11 that for every choice of positive semidefinite operators $`𝕂_1,\mathrm{},𝕂_d`$ of the form (7.14), satisfying the Stein identity (7.15), there exists a power series $`F𝒮𝒜_\mathrm{\Gamma }^{\mathrm{GR}}(𝒰,𝒴)`$ satisfying (besides (7.12)) supplementary interpolation conditions $`(X_LH_j)^L(T_L)\left[(X_LH_j)^L(T_L)\right]^{}`$ $`=𝕂_{j,L},`$ (7.16) $`(X_LH_j)^L(T_L)\left(G_jY_R\right)^R(T_R)`$ $`=𝕂_{j,LR},`$ $`\left[(G_jY_R)^R(T_R)\right]^{}(G_jY_R)^R(T_R)`$ $`=𝕂_{j,R}`$ for $`j=1,\mathrm{},d`$ and for some choice of associated functions $`H(z)`$ and $`G(z)`$ in representations (1.22), (1.24), (1.25) of $`F`$. Furthermore, all such $`F`$ can be parametrized by a linear fractional transformation. ###### Corollary 7.10. There is a power series $`F𝒮𝒜_\mathrm{\Gamma }^{\mathrm{GR}}(𝒰,𝒴)`$ satisfying the left interpolation condition $$\left(X_LF\right)^L(T_L)=Y_L$$ (7.17) if and only if there exist positive semidefinite operators $`𝕂_{1,L},\mathrm{},𝕂_{d,L}(𝒦_L)`$ that satisfy the Stein identity $$\underset{j=1}{\overset{d}{}}\left(𝕂_{j,L}N_j^{}𝕂_{j,L}N_j\right)=X_L^{}X_LY_L^{}Y_L.$$ (7.18) Again, for every choice of operators $`𝕂_{1,L},\mathrm{},𝕂_{d,L}`$ meeting conditions of Corollary 7.18, there exists $`F𝒮𝒜_\mathrm{\Gamma }^{\mathrm{GR}}(𝒰,𝒴)`$ satisfying (besides the left condition (7.17)) conditions (7.16) for $`j=1,\mathrm{},d`$ and for some choice of associated function $`H(z)`$ in representation (1.22) of $`F`$. ### 7.4. The Schur interpolation problem The classical Schur problem is concerned with necessary and sufficient conditions for existence of a (scalar valued) Schur function $`S`$ with the preassigned first $`n+1`$ Taylor coefficients at the origin (sometimes, especially if the Taylor coefficients at a point of $`𝔻`$ different from the origin, this problem is called the Carathéodory-Fejér problem). The operator-valued analogue of this problem is the following SP: given a collection of operators $`S_0,\mathrm{},S_n(𝒰,𝒴)`$, find necessary and sufficient conditions for existence of a Schur function $`S𝒮(𝒰,𝒴)`$ of the form $$S(z)=S_0+zS_1+\mathrm{}+z^{n1}S_{n1}+\mathrm{}.$$ The answer is given in terms of the Toeplitz matrix $$𝐒=\left[\begin{array}{cccc}S_0& 0& \mathrm{}& 0\\ S_1& S_0& \mathrm{}& \mathrm{}\\ \mathrm{}& \mathrm{}& \mathrm{}& 0\\ S_n& \mathrm{}& S_1& S_0\end{array}\right]$$ with operator entries: the SP has a solution if and only if the operator $`𝐒:𝒰^{n+1}𝒴^{n+1}`$ is contractive. An interpolation problem with the data string $`S_0,\mathrm{},S_N`$ containing gaps (that is, with unspecified $`S_k`$ for some $`k<n`$) also makes sense. In fact, this is a completion question: is it possible to complete a partially defined operator $`𝐒`$ as above to a contractive operator? This question (even in the scalar valued case) is beyond our current interests and will not be discussed here. Let $`\mathrm{\Gamma }`$ be an admissible graph and let $`_E`$ be the free semigroup generated by the edge set $`E`$ of $`\mathrm{\Gamma }`$. A subset $`_E`$ will be called lower inclusive if whenever $`v`$ and $`v=uw`$ for some $`u,w_E`$, then it is the case that also $`u`$. A natural noncommutative analogue of the Schur problem is the following: NSP: Let $`\mathrm{\Gamma }`$ be an admissible graph, let $`_E`$ be the free semigroup generated by the edge set $`E`$ of $`\mathrm{\Gamma }`$ and let $``$ be a finite lower inclusive subset of $`_E`$. Given a collection of operators $`\{S_v(𝒰,𝒴):v\}`$, find necessary and sufficient conditions for a noncommutative Schur-Agler function $$F(z)=\underset{v_E}{}F_vz^v𝒮𝒜_\mathrm{\Gamma }(𝒰,𝒴)$$ to exist such that $$F_v=S_v\text{for every}v.$$ (7.19) We will show that conditions (7.19) can be written in the form $$\left(X_LF\right)^L(T)=Y_L$$ (7.20) for an appropriate choice of $`X_L`$, $`Y_L`$ and $`T=\{T_e:eE\}`$; in other words we will show that the NSP is a particular left-sided case of Problem 1.8. The construction does not depend on the structure of the graph $`\mathrm{\Gamma }`$ and proceeds as follows. We are given a lower inclusive subset $``$ of the free semigroup $`_E`$ together with and operator $`F_v(𝒰,𝒴)`$ for each $`v`$. We let $`\mathrm{}^2()`$ be the Hilbert space with orthonormal basis $`\{\delta _v:v\}`$ indexed by $``$ and set $`𝒦_L=\mathrm{}^2()𝒴`$. Note that elements of $`𝒦_L`$ can also be viewed as functions $`vf(v)`$ on $``$ with values in $`𝒴`$ subject to $`_vf(v)_𝒴^2<\mathrm{}`$. Note that the empty word $`\mathrm{}`$ is in $``$ since $``$ is lower-inclusive. Define an operator $`X_L(𝒴,𝒦_L)`$ by $$X_L:y\delta _{\mathrm{}}y.$$ For each $`eE`$, we define an operator $`T_{L,e}`$ on $`𝒦_L`$ in terms of matrix entries $`T_{L,e}=[T_{L,e;v,w}]_{v,w}`$ (where each $`T_{L,e;v,w}(𝒴)`$) by $$T_{L,e;v,w}=\{\begin{array}{cc}I_𝒴\hfill & \text{if }v=we,\hfill \\ 0\hfill & \text{otherwise,}\hfill \end{array}$$ or via the equivalent functional form $$(T_ef)(v)=f(ve^1)\text{ for }f𝒦_L$$ where we use the convention (1.11) and we declare $`f(\text{undefined})=0`$. Then it is easily checked that, given a formal power series $`F(z)=_{v_E}F_vz^v`$, the left left evaluation with operator argument $`(X_LF)^L(T_L)`$ works out to be given by $$\left((X_LF)^L(T_L)u\right)(v)=F_vu\text{ for }v\text{ and all }u𝒰.$$ Hence, if we define $`Y_L:𝒰𝒦_L`$ by $$(Y_Lu)(v)=S_vu,$$ then the left tangential interpolation problem with operator argument associated with the data set $`𝒟=(T_L,X_L,Y_L)`$ is exactly equivalent to $`\mathrm{𝐍𝐒𝐏}`$, and hence necessary and sufficient conditions for the $`\mathrm{𝐍𝐒𝐏}`$ to have a solution can be derived from Theorem 7.2. ### 7.5. Interpolation with commutative data For this example we consider the general Problems 1.8 and 1.10 when the tuples $`T_L`$ and $`T_R`$ are commutative. As explained in Section 3, the interpolation conditions (1.3) imposed on a formal power series $`S𝒮𝒜_\mathrm{\Gamma }(𝒰,𝒴)`$ associated with Problem 1.3 can be expressed as interpolation conditions on the abelianized function $`S^𝐚`$ of commuting variables $`\lambda _{e_1},\mathrm{},\lambda _{e_d}`$: $$(X_LS^𝐚)^L(T_L)=Y_L,(S^𝐚Y_R)^R(T_R)=X_R.$$ (7.21) Similarly, the additional interpolation conditions (1.44)–(1.46) imposed on $`S𝒮𝒜_𝒢(𝒰,𝒴)`$ by Problem 1.10 can be expressed as interpolation conditions on the abelianized function $`S^𝐚`$: $`(X_LH_s^𝐚)^L(T_L)\left[(X_LH_s^{}^𝐚)^L(T_L)\right]^{}`$ $`=\mathrm{\Psi }_{s,s^{}}\text{ for }s,s^{}S:[s]=[s^{}],`$ $`(X_LH_s^𝐚)^L(T_L)\left(G_r^𝐚Y_R\right)^R(T_R)`$ $`=\mathrm{\Lambda }_{s,r}\text{ for }sS;rR:[s]=[r],`$ $`\left[(G_r^𝐚Y_R)^R(T_R)\right]^{}(G_r^{}^𝐚Y_R)^R(T_R)`$ $`=\mathrm{\Phi }_{r,r^{}}\text{ for }r,r^{}R:[r]=[r^{}].`$ (7.22) From the characterization of the class $`𝒮𝒜_𝒢(𝒰,𝒴)`$ as transfer functions of conservative SNMLSs with structure graph $`\mathrm{\Gamma }`$ and the counterpart of this result for the commutative Schur-Agler class $`𝒮𝒜_{Z_\mathrm{\Gamma }^𝐚}(𝒰,𝒴)`$ found in , it is clear that the abelianization $`S^𝐚`$ of any element $`S𝒮𝒜_\mathrm{\Gamma }(𝒰,𝒴)`$ is an element of $`𝒮𝒜_{Z^𝐚}(𝒰,𝒴)`$ as studied in , and, conversely, any element $`F`$ of $`𝒮𝒜_{Z_\mathrm{\Gamma }^𝐚}(𝒰,𝒴)`$ lifts to an element $`S𝒮A_\mathrm{\Gamma }(𝒰,𝒴)`$ (so $`F=S^𝐚`$). The results of can be applied to the abelianized problems involving interpolation conditions (7.21) (and possibly also (7.22)) for a function $`S^𝐚`$ in the commutative Schur-Agler class $`𝒮𝒜_{Z_𝒢^𝐚}(𝒰,𝒴)`$. When this is done, the Stein equation (1.39) is the same as the Stein equation in where it was shown to be the necessary and sufficient condition for the abelianized interpolation problem to have a solution in the commutative Schur-Agler class $`𝒮𝒜_{Z_\mathrm{\Gamma }^𝐚}(𝒰,𝒴)`$. In this way, we see that interpolation problems for formal power series in noncommuting indeterminants involving commutative data reduces to the more standard interpolation problems for analytic functions in commuting variables. As an example, let us consider the case with commutative data for the noncommutative-ball setting discussed in Section 7.2. Let, in particular, $`𝒦_L=^n`$, let $`T_L`$ is the $`d`$-tuple of diagonal matrices constructed from $`n`$ points $`\lambda _i=(\lambda _i^{(1)},\mathrm{},\lambda _i^{(d)})𝔹^d`$ ($`i=1,\mathrm{},n`$) by $$T_{L,j}=\mathrm{diag}(\lambda _1^{(j)},\mathrm{},\lambda _n^{(j)})\text{for}j=1,\mathrm{},d,$$ and let $`X_L`$ and $`Y_L`$ be conformally decomposed as $$X_L=\left[\begin{array}{c}b_1\\ \mathrm{}\\ b_n\end{array}\right]\text{and}Y_L=\left[\begin{array}{c}c_1\\ \mathrm{}\\ c_n\end{array}\right].$$ Then the pair $`(T_L,X_L)`$ is left admissible and it is easily seen that $$(X_LF)^L(T_L)=\mathrm{Col}_{1in}b_iF(\lambda _i)$$ (where $`F(\lambda _i)`$ is defined via (3.6), so that condition (7.1) collapses to $`n`$ left sided conditions $$b_iF(\lambda _i)=c_i(i=1,\mathrm{},n).$$ (7.23) Furthermore, the matrix $`𝕂_L`$ in (7.5) admits a more explicit representation $$𝕂_L=\left[\frac{b_ib_j^{}c_ic_j^{}}{1\lambda _i,\lambda _j}\right]_{i,j=1}^n$$ (7.24) where $`\lambda _i,\lambda _j`$ stands for the standard inner product in $`^d`$. Thus, Theorem 7.4 gives \[39, Theorem 4.1\]: there exists a power series $`F𝒮𝒜_{\mathrm{\Gamma }^{\mathrm{FM}}}(𝒰,𝒴)`$ satisfying interpolation conditions (7.23) if and only if the matrix $`𝕂_L`$ defined in (7.24) is positive semidefinite. We note that the commutative (several-variable) analogue of the Schur problem discussed above in Section 7.4 is one of the examples for the commutative theory discussed in .
warning/0506/gr-qc0506111.html
ar5iv
text
# A non-singular black hole model as a possible end-product of gravitational collapse. ## 1 Introduction It is generally accepted that the formation of a black hole proceeds with the formation of a marginally trapped surface (the apparent horizon) into which matter is collapsing and one that encloses a region of space interior to which the light cones are tipped so much so that even outgoing null geodesics proceed toward ever decreasing values of the radial coordinate. The famous singularity theorems , then posit that under certain assumptions on energy conditions of matter, the eventual state of the collapse is an infinite density singularity surrounded by an event horizon (defined by the outermost trapped surface). It is, however, a fact that the existence of such a singularity is neither easy to verify nor refute observationally because of the limits set by the black hole event horizon. In the absence of such observational evidence, the compelling logic in the arguments leading to the singularity theorems has generally shaped the widespread conviction that indeed a black hole must contain a physical singularity. As is known, however, the picture of a singular black hole opens up several puzzles. For example, the existence of spacetime singularities results in irretrievably total information loss . Moreover, the infinite tidal forces that result from such a singularity formation lead to a breakdown in the descriptive power of General Relativity . As a result it is generally believed that the perceived final collapse state implied by the singularity theorems could be but a manifestation of the incompleteness of classical gravity and that a correct (quantum) theory of gravity will dispense with this singular final state and admit a state of finite (albeit very high) curvature. An altogether different view is that the singularities implied by the aforementioned theorems simply reflect our lack of knowledge of the properties of matter under extreme conditions. Thus it has been suggested that the singularity theorems could be circumvented if (some of) the conditions imposed on the properties of collapsing matter were relaxed. Indeed modern frameworks (like string theory) attempting to formulate a quantum theory of gravity seek the existence of a fundamental length scale and hence a singularity free spacetime. There are also philosophical questions as to whether spacetime can indeed accommodate singularities of any kind, be they from gravitational collapse or vacuum-induced . As a result of these and other considerations a renewed interest has grown lately with regard to spacetime singularities and their existence . Indeed the concept of non-singular collapse dates back to Sakharov’s consideration of the equation of state $`p=\rho `$ for a superdense fluid and Gliner’s view that such a fluid could be the final state of gravitational collapse. Such ideas are now attracting a growing amount of attention. For example, Ellis has argued that in the case of a closed universe a trapped surface does not necessarily lead to a singularity. Recently the present authors have investigated an analogous argument with regard to black holes in . It was pointed out in that under physically acceptable energy conditions the expansion of the outgoing mode which assumes increasingly negative values in a trapped region can eventually turn around to (even) assume a positive signature inside a black hole. Such behavior suggests the existence some interior region, surrounding the core, which is not trapped. In turn, this feature implies the absence of a singularity in such a spacetime. In recent years, several non-singular black hole solutions have been found. Dymnikova, for example, has constructed an exact solution of the Einstein field equations for a non-singular black hole containing at the core a fluid de-Sitter-like fluid with anisotropic pressure. This solution is Petrov Type \[(II),(II)\] by classification<sup>1</sup><sup>1</sup>1We shall present a solution with the same asymptotic behavior as in which however is different in both classification (it is Petrov Type \[II,(II)\] and physical interpretation.. Another line of investigation in this area, initiated by Markov , suggests a limiting curvature approach. This idea has been explored further by several authors, see for example . A common feature in all these treatments is that the geometry of the spacetime in question is Schwarschild at large $`r`$ as expected by Birkoff’s theorem, and de Sitter-like at small $`r`$ values as implied by the $`p=\rho `$ equation of state assumed to prevail near $`r=0`$. These approaches can be put in two broad classes. (a) Those for which the transition to this “exotic” state of matter can be placed well inside the Schwarzschild horizon and (b) those which replace the entire volume of the Schwarschild metric interior to $`r=2M`$ by a substance with the $`p=\rho `$ equation of state , dispensing in this way with the presence of the horizon altogether and the “information paradox” problems it engenders. The problem of transition from the usual matter equation of state to that appropriate for these non-singular solutions is usually not addressed at all. This problem is particularly acute for models of the second class discussed above, since the density at horizon formation scales like $`\rho 10^{16}(M/M_{})^2`$ g cm<sup>-3</sup> and for objects of mass $`M10^8M_{})`$, is not greater than that of water, which is known to have an equation of state very different from $`p=\rho `$. Moreover, issues to do with direct matching of an external Schwarschild vacuum to an interior de Sitter have previously been discussed based on junction conditions . In fact, in terms of gravitational collapse the associated difficulties have been used to suggest modifications in some treatments . The present work is motivated by these questions and is undertaken as a first effort to provide some answers by constructing models using an explicit equation of state with the desired properties. It would appear that the junction constraints when applied to the end-product of non-singular gravitational collapse imply the matter fields across the Schwarschild/de Sitter boundary should have a radially dependent equation of state of the form $`1w(r)1`$ that smoothly changes the matter-energy from a stiff fluid to a cosmological constant. The model sketched in this paper describes the geometry of a body that passed during its collapse through a stiff fluid state to settle into the final state of a non-singular black hole. The solution allows for a matter fluid region (with a family of equations of state $`0w\left(r\right)1`$ which envelopes an intermediate region with a family of equations of state $`1<w\left(r\right)<0`$ which in turn envelopes a de Sitter like region with $`w=1`$ at the center. Our model differs both from the traditional singular black hole solutions which introduce matter at the singularity and from the relatively new non-singular solutions which usually introduced only a de sitter-like spacetime inside the black hole. The main feature of our model is that it (1) introduces gravitating matter inside a non-singular black hole and (2) offers a reasonable explanation of how part of this matter can evolve towards a de Sitter-like vacuum to provide the radial tension or negative pressure that supports the remaining matter fields against the formation of a singularity. The rest of the paper is organized as follows. In section 2 we summarize the equations to be solved. We also construct the working equation of state and discuss some of its desirable features. In section 3 we solve the Einstein Field Equations for the desired geometry and obtain an exact solution. In section 4 we highlight the physical implications of the model. Section 5 concludes the discussion. ## 2 Problem formulation ### 2.1 The field equations It is assumed, for simplicity, that the collapsed object can be reasonably described by a spherical, static geometry. The desired line element then takes the general form $$ds^2=e^{\nu \left(r\right)}dt^2+e^{\lambda \left(r\right)}dr^2+r^2\left(d\theta ^2+\mathrm{sin}^2\theta d\phi ^2\right),$$ (2.1) where $`\lambda \left(r\right)`$ and $`\nu \left(r\right)`$ are to be determined from the Einstein Field Equations $$G_{\mu \nu }=8\pi GT_{\mu \nu }.$$ (2.2) Under the assumed spacetime symmetry, Eqs.2 reduce to $$e^\lambda \left(\frac{\lambda ^{}}{r}\frac{1}{r^2}\right)+\frac{1}{r^2}=8\pi G\rho \left(r\right),$$ (2.3) $$e^\lambda \left(\frac{\nu ^{}}{r}+\frac{1}{r^2}\right)\frac{1}{r^2}=8\pi Gp_r\left(r\right),$$ (2.4) and $$e^\lambda \left(\frac{\nu ^{\prime \prime }}{2}\frac{\lambda ^{}\nu ^{}}{4}+\frac{\nu ^2}{4}+\frac{\nu ^{}\lambda ^{}}{2r}\right)=8\pi p_{}\left(r\right).$$ (2.5) Here $`\rho \left(r\right)=`$ $`T_0^0`$ is the energy density, $`p_r=T_1^1`$ is the radial pressure and $`\left(p_\theta =T_2^2\right)=\left(p_\phi =T_3^3\right)=p_{}`$ is the tangential pressure. In this treatment the fluid is anisotropic with $`T_2^2=T_3^3T_1^1`$ and $`T_1^1T_0^0T_2^2`$. The spacetime is thus Petrov Type \[II, (II)\], where ( ) implies a degeneracy in the eigenvalues of the Weyl tensor. ### 2.2 Equation of state In the remaining part of this section we construct the working equation of state $`p_r\left(\rho \right)`$ in the radial direction which reproduces the characteristics of the collapsed body as highlighted in the previous section. We justify the choice by pointing out its desired features. On the other hand, the tangential pressure equation of state $`p_{}\left(\rho \right)`$ will be derived later from the Einstein Field Equations (Eq. 2.5) using the radial pressure function $`p_r\left(\rho \right)`$ to be constructed. The main desired features are that (1) the equation of $`p_r\left(\rho \right)`$ change from the one associated with the usual matter, at low densities, to that associated with the de Sitter geometry at sufficiently high densities near the center of the configuration and; (2) that this change be well-behaved. To this end we consider an equation of state that takes the general form<sup>2</sup><sup>2</sup>2This choice may not be unique. where $`m`$ and $`1/n`$ are, yet to be determined, positive real numbers and $`\alpha `$ is a free parameter to be constrained. This equation implies presence of a $$p_r\left(\rho \right)=\left[\alpha \left(\alpha +1\right)\left(\frac{\rho }{\rho _{\mathrm{max}}}\right)^m\right]\left(\frac{\rho }{\rho _{\mathrm{max}}}\right)^{1/n}\rho ,$$ (2.6) maximum limiting density $`\rho _{\mathrm{max}}`$ concentrated in a region of size $`r_0=\sqrt{\frac{1}{G\rho _{\mathrm{max}}}}`$, the core region of the configuration. One notes that the main desirable features are readily apparent in the functional form of Eq. 2.6. At low densities, $`\left(\frac{\rho }{\rho _{\mathrm{max}}}\right)^m<\frac{\alpha }{\alpha +1}`$, the equation of state reduces to that of a polytrope of index $`n`$ ($`p_r\rho ^{1+1/n}`$) while for $`\left(\frac{\rho }{\rho _{\mathrm{max}}}\right)^m>\frac{\alpha }{\alpha +1}`$ the pressure decreases to eventually approach that of the vacuum $`p_r=\rho _{\mathrm{max}}`$ as $`\rho \rho _{\mathrm{max}}`$. We look for the simplest form of Eq. 2.6 by judiciously constraining both the indices $`m`$ and $`n`$ and the parameter $`\alpha `$. Since Eq. 6 already satisfies the general features desired by the model, its specific form, based on an appropriate choice of the indices $`m`$ and $`n`$ and the parameter $`\alpha `$ is made through the demand that it must satisfy the following basic conditions. (1) It must have no pathologies, i.e. (i) the sound speed $`\frac{dp}{d\rho }`$ can not be a maximum at $`\rho =0`$ (ii) in order to rule out superluminal behavior, $`\frac{dp}{d\rho }1`$, the maximum sound speed, i.e. the value of $`\frac{dp}{d\rho }`$ at the point where $`\frac{d^2p}{d\rho ^2}=0`$ must be given by $`\frac{dp}{d\rho }_{\rho _{stiff}}=c_s^2=c=1`$. (2) It must satisfy minimum acceptable energy conditions, namely the Weak Energy Condition, $`\rho 0`$, $`\rho +p_r0`$ and the Dominant energy Condition, $`\rho 0`$, $`p_r[\rho ,^+\rho ]`$. The $`\left[m=1,1/n=0\right]`$ and the $`\left[m=2,1/n=0\right]`$ cases are both pathological (they do not satisfy (i) above) and will therefore be discarded. The next simplest potential choices are $`\left[m=1,n=1\right]`$ and $`\left[m=2,n=1\right]`$. Both these forms, apparently, manifest no pathologies of the previous cases. In both cases the sound speed vanishes $`c_s0`$ at vanishing density $`\rho 0`$. Furthermore for $`\rho >0`$ the sound speed $`c_s`$ is initially an increasing function of $`\rho `$, as expected in reality??. One can therefore demand in each case (i.e. for fixed $`m`$ and $`n`$ values ) that the maximum sound speed coincide with that of light i.e. the extremum sound speed at $`\frac{d^2p}{d\rho ^2}=0`$ be given by $`\frac{dp}{d\rho }=1`$ in order to rule out superluminal behavior. These conditions also suffice to determine the parameter $`\alpha `$ and allow the fulfillment of the energy conditions in (2). Solving $`\frac{d^2p}{d\rho ^2}=0`$ and $`\frac{dp}{d\rho }=1`$ simultaneously for the $`\left[m=1,n=1\right]`$ case, i.e. fixes $`\alpha `$ at $`\frac{3}{4}\frac{1}{4}\sqrt{33}`$ or $`\frac{1}{4}\sqrt{33}+\frac{3}{4}`$ (we keep only the positive solution). Similarly for the $`\left[m=2,n=1\right]`$ one fixes $`\alpha `$ to $`\alpha =\mathrm{2.\hspace{0.17em}213\hspace{0.17em}5}`$. We will select the quartic function corresponding to $`m=2,n=1`$ as the representative form for our model equation of state, in part because, unlike the cubic form, it unambiguously provides only one possible value of $`\alpha `$ and makes the interpretation simpler. Thus Eq. 2.6 now takes the form $$p_r\left(\rho \right)=\left[\alpha \left(\alpha +1\right)\left(\frac{\rho }{\rho _{\mathrm{max}}}\right)^2\right]\left(\frac{\rho }{\rho _{\mathrm{max}}}\right)\rho ,$$ (2.7) with the constraint that $`\alpha =\mathrm{2.\hspace{0.17em}213\hspace{0.17em}5}`$. It is the equation of state we adapt throughout the remaining part of the paper. A few words on the behavior of our assumed equation of state. While no explicit microscopic prescription leading to its form is presently given, we consider that its softening past the “stiff” ($`c_s^2=dp/d\rho =1`$) state and its eventual conversion to an equation of state appropriate to that of the vacuum ($`p=\rho `$), is effected by the coupling of matter to a scalar field akin to that of Higgs that gives rise to particle masses. The dominance of the pressure by terms which render it negative is considered to imply that eventually the energy associated with the self interaction of this field provides the dominant contribution to the energy momentum tensor, which is assumed to be that of a perfect fluid. In this paper no attempt is made to follow the dynamical evolution of the collapse. Instead it is assumed that the collapsed body has already reached static equilibrium from its collapse. Thus, we only investigate the characteristics of the various static 2-surfaces as the radial coordinate decreases from the matter surface $`r=R`$ to the body center $`r=0`$. Since our anticipated solution will allow matter fields we can, with no loss of generality, take for initial conditions on the collapsed body surface $`R`$ to be $`\rho =0,p_r=0`$. We therefore expect to have four regions in this spacetime which must be satisfied by the expected solution. Region I: Schwarzschild vacuum: $`R<r<\mathrm{}`$, $`p=\rho =0`$ Region II: Regular matter fields: $`r_\epsilon <r<R,\rho >0,0<p\rho ,`$ Region III: Quintessential fields: $`r_0<r<r_\epsilon ,0<\rho <\rho _{\mathrm{max}},\rho _{\mathrm{max}}<p<0,`$ Region IV: $`\mathrm{\Lambda }`$vacuum: $`0rr_0,p=\rho =\rho _{\mathrm{max}}.`$ The solution to Eqs. 2.1-2.5, must satisfy the following asymptotic conditions at large $`r`$ and small $`r`$, respectively. (i) The spacetime must be asymptotically Schwarzschild for large $`r`$, i.e. for $`R<r<\mathrm{}`$ $$ds^2=\left(1\frac{2M}{r}\right)dt^2+\frac{1}{\left(1\frac{2M}{r}\right)}dr^2r^2\left(d\theta +\mathrm{sin}^2\theta d\phi ^2\right).$$ (2.8) where for the black hole $`R`$ is some hypersurface such that $`R<2M`$, $`M=4\pi _0^{\mathrm{}}\rho \left(r\right)r^2𝑑r`$being the total mass. (ii) The spacetime must be asymptotically de Sitter $$ds^2=\left(1\frac{r^2}{r_0^2}\right)dt^2+\frac{1}{\left(1\frac{r^2}{r_0^2}\right)}dr^2+r^2\left(d\theta +\mathrm{sin}^2\theta d\phi ^2\right).$$ (2.9) for $`0r\left(r_0<R\right)`$. Here, $`r_0=\sqrt{\frac{3}{8\pi G\rho _{\mathrm{max}}}}=\sqrt{\frac{3}{\mathrm{\Lambda }}}`$ signals the onset of de Sitter behavior, where $`\rho _{\mathrm{max}}=\rho _{r0}`$ is the upper-bound on the density of the fields of order of Planck density $`\rho _{Pl}`$. The asymptotic conditions in Eqs. 2.8 and 2.9 imply there is an interior region which includes the family of surfaces $`\mathrm{\Sigma }=\left\{\mathrm{\Sigma }_{II}\mathrm{\Sigma }_{III}\right\}`$ and which interfaces with region I on the outer side and region IV on the inner side. The entire spacetime must therefore satisfy regularity conditions at the two interfaces $`r=R`$ and $`r=r_0`$. Put simply, such conditions guarantee (i) continuity of the mass function and (ii) continuity of the pressure across the interfacing hyperfaces. Thus at each interface, i.e. $`(I,II)`$ and $`(III,IV)`$ we must have $`\left[\rho ^+\rho ^{}\right]_{r=r_i}`$ $`=0,`$ $`\left[p^+p^{}\right]_{r=r_i}`$ $`=0,`$ where $`r_i=\{R,r_0\}`$ and $`+`$, $``$ refer to the exterior and interior values respectively and $`p=\{p_r,p_{}\}`$. A desirable feature of the model (as we find later) is the smooth continuity of the density and pressure between region II and region III. This feature removes the problem having to match the two regions through junction conditions between the regular matter fields and the vacuum-like field, since here such conditions will be satisfied naturally. ## 3 The solution We now solve the Einstein equations 2.3-2.5 for the spacetime of the model. Integration of Eq. 2.3 gives $$e^\lambda =18\pi \frac{1}{r}_0^r\rho \left(r^{}\right)r^2𝑑r^{}=1\frac{2m\left(r\right)}{r},$$ (3.1) where, as stated before, $`m\left(r\right)`$ is the mass enclosed by the a 2-sphere of radius $`r`$. Further, using Eqs. 2.7 and 11, integration of the $`T_1^1`$ Eq. 2.4, gives $$e^\nu =\left(1\frac{2m\left(r\right)}{r}\right)e^{{\scriptscriptstyle 8\pi \left[\alpha \left(\alpha +1\right)\left({\scriptscriptstyle \frac{\rho }{\rho _{\mathrm{max}}}}\right)^2\right]\left({\scriptscriptstyle \frac{\rho }{\rho _{\mathrm{max}}}}\right)\rho {\scriptscriptstyle \frac{dr}{1\frac{2m\left(r\right)}{r}}}}},$$ (3.2) where $`\alpha =\mathrm{2.\hspace{0.17em}213\hspace{0.17em}5}`$. Eqs. 3.1 and 3.2 into Eq. 2.1 give $$ds^2=\left(1\frac{2m\left(r\right)}{r}\right)e^{\mathrm{\Gamma }\left(r\right)}dt^2+\frac{1}{1\frac{2m\left(r\right)}{r}}dr^2+r^2\left(d\theta ^2+\mathrm{sin}^2\theta d\phi ^2\right),$$ (3.3) where $$\mathrm{\Gamma }\left(r\right)=8\pi \left[\alpha \left(\alpha +1\right)\left(\frac{\rho }{\rho _{\mathrm{max}}}\right)^2\right]\left(\frac{\rho }{\rho _{\mathrm{max}}}\right)\left(\frac{r}{r2m\left(r\right)}\right)\rho 𝑑r$$ (3.4) Eq. 3.3 is the general solution for the geometry of our model. It describes the spacetime of a non-singular black hole with both matter and a de Sitter core. We shall describe below a particular solution resulting from the choice of a prescribed density function $`\rho \left(r\right)`$. To formally complete the solution one must also solve equation 5 for the tangential pressure $`p_{}=\{p_\theta ,p_\phi \}`$. On doing this one finds (see also ) that $$p_{}=p_r+\frac{r}{2}p_r^{}+\frac{1}{2}\left(p_r+\rho \right)\left[\frac{Gm\left(r\right)+4\pi Gr^3p_r}{r2Gm\left(r\right)}\right],$$ (3.5) where $`p_r`$ is given by Eq. 2.7. Note that in this model the last term in Eq. 3.6 is not vanishing, in general, as is the case considered in some previous treatments where it was assumed that $`p_r=\rho `$ for all $`\rho `$ (see for example ). Eq.3.6 is a generalization of the Tolman-Oppenheimer-Volkoff equation . One can now discuss the results of Eqs.3.1-3.4. The mass $`m\left(r\right)`$ enclosed a 2-surface at any radial coordinate $`0<r<R`$ is given by $`m\left(r\right)=4\pi _0^r\rho \left(r^{}\right)r^2𝑑r^{}`$. The entire mass of the body $`M`$ is given by $`M=_0^{\mathrm{}}m\left(r\right)𝑑r=_0^{\mathrm{}}\rho \left(r\right)r^2𝑑r`$. It is then clear that outside the mass of the body $`T_0^0=0`$. Further, as $`\rho `$ vanishes outside the body, so does the pressure (see Eqs. 2.7 and 3.3). In particular, from Eq. 3.3 one observes that in the limit $`\rho 0`$ and $`m\left(r\right)M`$ the lapse function and the shift vector give $`e^\nu =e^\lambda =1\frac{2GM}{r}`$. Thus outside the mass (i.e. for large $`r`$) the spacetime becomes asymptotically Schwarzschild, with the metric given by Eq. 2.8. Further, in the limit $`rr_0`$, $`\rho \rho _{\mathrm{max}}`$ (and as both Eqs. 2.7 and 3.4 show) $`p_r\rho _{\mathrm{max}}`$ and $`p_{}\rho _{\mathrm{max}}`$. It follows, therefore that for $`0r<r_0`$ the fluid has constant positive density $`\rho _{\mathrm{max}}`$ and constant negative pressure $`\rho _{\mathrm{max}}`$ and takes on the character of a de Sitter spacetime, through the equation of state $`p=\rho _{\mathrm{max}}`$. In order to establish the actual asymptotic behavior of the spacetime at small $`r`$ one has to choose a density profile for the matter. We choose to use the one employed by Dymnikova because of its simplicity and convenient form for integration over the source volume $`\rho =\rho _{\mathrm{max}}\mathrm{exp}\left[\frac{r^3}{r_g\left(r_0\right)^2}\right]`$. The mass $`m\left(r\right)`$ enclosed by a 2-sphere at the radial coordinate $`r`$ is then given by $`m\left(r\right)=_0^r\rho _{\mathrm{max}}\{\mathrm{exp}\left[\frac{r^3}{r_g\left(r_0\right)^2}\right]\}r^2𝑑r^{}=M\left[1\mathrm{exp}\left(\frac{r^3}{r_g\left(r_0\right)^2}\right)\right]`$, where $`M`$ is the total mass $`M=_0^{\mathrm{}}\rho _{\mathrm{max}}\{\mathrm{exp}\left[\frac{r^3}{r_g\left(r_0\right)^2}\right]\}r^2𝑑r`$. The entire mass is essentially concentrated in a region of size $`R(r_0^2r_g)^{1/3}`$, which in general is close to $`r=0`$, with its precise value depending on the value of the maximum density assumed $`\rho _{\mathrm{max}}`$. The introduction of the density function $`\rho =\rho _{\mathrm{max}}\mathrm{exp}\left[\frac{r^3}{r_g\left(r_0\right)^2}\right]`$ in Eq. 3.3 provides a particular solution for the model. In the limit $`r0`$, it is seen that the solution Eq. 3.3 asymptotically leads to Eq. 2.9 as the spacetime becomes asymptotically de Sitter. ## 4 Physical Interpretation Eqs. 3.3, 3.4, 3.5 along with the prescribed density profile, form the solution in this model describing the spacetime of a non-singular black hole containing both matter and a de Sitter core. A heuristic interpretation of the form of the equation of state used has been given in the previous section. The field whose presence leads to the asymptotic relation $`p=\rho `$ has not been specified but it must be similar to the quintessence field used in cosmology. In this specific case, it could in fact be the Higgs field that gives rise to the particle masses. As such the maximum density could be roughly $`\rho _{\mathrm{max}}m_H^4`$, the latter being the Higgs field expectation value. This could be either of order 1 TeV or even of order of the Plank mass if, as argued, that should be the typical mass of all scalars coupled to gravity. The depicted scenario allows two viable, albeit, different interpretations in describing the internal structure of such a black hole. The two interpretations have the common feature that in both cases a fluid with a negative pressure, in the interior, supports the matter fields in the outer region against any further collapse. The effect in both cases is to create a non-singular black hole. ### 4.1 The quintessential picture The first interpretation is directly based on the analytical functional form of the equation of state depicted in Eq. 2.7. Here, as one moves from the surface $`R`$ of the collapsed body towards the center, one first encounters region II where from the functional form of the equation of state one crosses a family $`\mathrm{\Sigma }_{II}`$ of 2-surfaces with matter-like $`w>0`$. At the critical point $`\rho _c`$ in the $`\rho p_r`$ space $`\frac{dp_r}{d\rho }=0`$ thereafter the pressure $`p_r`$ begins to drop with increasing density $`\rho `$. At some point $`\rho _c=\left(\frac{\alpha }{\left(\alpha +1\right)}\right)^{\frac{1}{2}}\rho _{\mathrm{max}}`$, the fluid temporarily becomes pressureless $`w0`$, implying that $`T_1^1=G_1^10`$. Beyond this, for smaller values of $`r`$ one enters region III in which one is now crossing a family of 2-surfaces $`\mathrm{\Sigma }_{III}`$ for which the equation of state is given by $`1<w<0`$, smoothly decreasing from $`w=0`$ towards $`w=1`$. In this picture we will refer to this as the quintessential region. Evidently $`\mathrm{\Sigma }_{III}`$ exhausts the entire family of quintessential fluids $`1<w<0`$. This fluid family provides some of the negative pressure against the implosion of the matter fields in the outer region. The rest of the negative pressure is provided by the constant density $`\rho _{\mathrm{max}}`$, constant pressure $`p=\rho _{\mathrm{max}}`$, inner core $`0<r<r_0`$ that mimics the cosmological constant. This interpretation is valid provided the equation of state $`p_r\left(\rho \right)`$ is, indeed in reality, a well-behaved function of $`r`$ for $`\rho _c<\rho <\rho _{\mathrm{max}}`$. In this quintessential picture, one can estimate the amount remaining pure matter (the total field for which $`w0`$) as a fraction of the total energy of the system. This matter lies in the outer region $`r_\epsilon rR`$ (i.e. between points B and C in Figure 1). Since B is given by, $`\frac{dp_r}{d\rho }_{\rho _c}=0`$ and C is given by $`p_r_{\rho _\epsilon }=0`$, one can infer from Eq. 7 that in this region the density function $`\rho `$ is bounded by $`\left[\rho _c=\left(\frac{\alpha }{3\left(\alpha +1\right)}\right)^{\frac{1}{2}}\rho _{\mathrm{max}}\right]\rho \left[\rho _\epsilon =\left(\frac{\alpha }{\left(\alpha +1\right)}\right)^{\frac{1}{2}}\rho _{\mathrm{max}}\right]`$. To determine the matter mass <sub>Mmatter</sub> enclosed we use the density profile $`\rho \left(r\right)=\rho _{\mathrm{max}}\mathrm{exp}\left[\frac{r^3}{r_0^2r_g}\right]`$ and perform the integral $`\rho r^2𝑑r`$ with the appropriate limits. It is useful to change the independent variable from $`r`$ to $`\rho `$. Then $`r^2dr=\frac{1}{3}r_0^2r_g\left(\frac{\rho _m}{\rho }\right)d\left(\frac{\rho }{\rho _m}\right)`$ and we have that $`{}_{matter}{}^{}=4\pi \rho r^2dr=\frac{4}{3}\pi r_0^2r_g_0^{\rho _\epsilon }d\rho `$. Thus $$_{matter}=\frac{4}{3}\pi r_0^2r_g\left(\frac{\alpha }{\alpha +1}\right)^{\frac{1}{2}}\rho _{\mathrm{max}}.$$ (4.1) Using $`\alpha =\mathrm{2.\hspace{0.17em}213\hspace{0.17em}5}`$ as adopted earlier for our model, gives $`M_{matter}=\mathrm{0.829\hspace{0.17em}95}\left(\frac{4}{3}\pi r_0^2r_g\right)\rho _{\mathrm{max}}`$. On the other hand the total mass $`M=4\pi \rho \mathrm{exp}\left[\frac{r^3}{r_0^2r_g}\right]r^2𝑑r=\frac{4}{3}\pi r_0^2r_g\rho _{\mathrm{max}}`$. Thus the fractional content of matter in the quintessential picture for the black hole is $`\frac{_{matter}}{M}=\mathrm{0.829\hspace{0.17em}95}0.83`$. ### 4.2 A two-fluid system One can attempt yet a seemingly equally viable interpretation of the black hole’s internal structure. If one holds as in the view of a phase transition of the fluid from the matter-like form to the de Sitter state given by $`p=\rho _{\mathrm{max}}`$, then the successive 2-surfaces in the family $`r_0<r<r_c`$ will represent a two fluid system of matter fields and the $`\mathrm{\Lambda }`$vacuum which in both dynamic and thermodynamic equilibrium. The fluid begins as pure matter on the 2-surface $`r_c`$ and gets richer and richer in the $`\mathrm{\Lambda }`$vacuum state as one moves deeper and deeper into the black hole. Then the density $`\rho _r\left(r\right)`$ at any such point on any such 2-surface shell $`r_0<r<R`$ is really a sum of two partial densities $$\rho =\beta \left(r\right)\rho _c\left[1\beta \left(r\right)\right]\rho _{\mathrm{max}},$$ (4.2) where $`\rho _c`$ is the critical density just before collapse when $`p=\rho _c`$ and $`0\beta 1`$. The function $`\beta \left(r\right)`$ is the fractional content of pure matter fields to the total surface energy contained in an elementary 2-surface shell $`4\pi r^2dr`$ at $`r_0<r<R`$. For decreasing $`r`$, $`\beta \left(r\right)`$ is a decreasing function, evolving from $`\beta \left(R\right)=1`$ to $`\beta \left(r_0\right)=0`$. In terms of the partial pressures of the matter fields and the de Sitter vacuum, the pressure at any position $`r_0r<R`$ is then given by $$p_r\left(r\right)=\underset{n=0}{\overset{3}{}}a_n\left(\frac{\rho _c}{\rho _{\mathrm{max}}}\right)^n\left[\beta \left(r\right)\rho _c\left(1\beta \left(r\right)\right)\rho _{\mathrm{max}}\right]$$ (4.2) where the $`a`$’s are positive constants given respectively by $`a_0`$ $`=\left(1\beta \right)\left[\alpha \left(\alpha +1\right)\left(1\beta \right)^2\right],`$ $`a_1`$ $`=\beta \left[\alpha +\left(\alpha +1\right)\left(1\beta \right)\right],`$ $`a_2`$ $`=\beta ^2\left(\alpha +1\right)\left(1\beta \right),`$ (4.4) $`a_3`$ $`=\beta ^3\left(\alpha +1\right).`$ This expression based on partial pressures will then replace Eq. 2.7 as the equation of state $`p_r\left(\rho \right)`$ in the event the evolution of matter fields to a de Sitter state involves a spontaneous phase transition. One notes that the pressure $`p_r`$ still has the desired asymptotic behavior. Thus in the limit $`\beta \left(r\right)1`$, $`pp_c=\left[\alpha \left(\alpha +1\right)\left(\frac{\rho _c}{\rho _{\mathrm{max}}}\right)^2\right]\left(\frac{\rho _c}{\rho _{\mathrm{max}}}\right)\rho _c`$, while in the limit $`\beta \left(r\right)0`$, $`p\rho _{\mathrm{max}}`$ as expected. One can now estimate the energy $`M_{matter}`$ in the pure matter fields as a fraction of the total energy of the black hole in this two-fluid picture. We assume that up to the critical density $`\rho _c`$ the matter is pure and that the energy content contribution of the de Sitter fluid with an energy density $`\rho _{\mathrm{max}}`$ starts to grow at $`\rho =\rho _c`$. Then the matter content can be written as a two piece integral $`M_{matter}=4\pi _{r_c}^{\mathrm{}}\rho r^2𝑑r+4\pi _{r_0}^{r_c}\beta \left(r\right)\rho _cr^2𝑑r`$. We will assume $`\beta \left(r\right)`$ to be simply linear in $`r`$ with a form $`\beta \left(r\right)=\frac{rr_0}{r_cr_0}`$, which satisfies the requirements imposed on it above. Then substituting for $`\beta \left(r\right)`$ and again noting that $`r^2dr=\frac{1}{3}r_0^2r_g\left(\frac{\rho _m}{\rho }\right)d\left(\frac{\rho }{\rho _m}\right)`$ we get $$_{matter}=\frac{4}{3}\pi \left\{\left[r_0^2r_g\right]+\left[\frac{1}{4}\left(r_c+r_0\right)\left(r_c^2+r_0^2\right)\frac{1}{3}r_0\frac{r_c^3r_0^3}{r_cr_0}\right]\right\}\rho _c.$$ (4.5) From the definition of the density function we have that $`r=\left[r_0^2r_g\mathrm{ln}\left(\frac{\rho }{\rho \mathrm{max}}\right)\right]^{\frac{1}{3}}`$. Since $`rr_0`$ as $`\rho \rho _{\mathrm{max}}`$ we have that for $`0<r<r_0`$, $`r\left(\rho \right)`$ is not well defined since $`r`$ is now not single-valued in $`\rho `$. Further, we don’t know the actual size of $`r_0`$. As a result we shall assume, for the purpose of evaluating Eq. 4.5, that both $`r=\left[r_0^2r_g\mathrm{ln}\left(\frac{\rho }{\rho \mathrm{max}}\right)\right]^{\frac{1}{3}}`$ and $`rr_0`$ as $`\rho \rho _{\mathrm{max}}`$ hold so that $`r_0\left(\rho _{\mathrm{max}}\right)0`$. This is equivalent to having the de Sitter vacuum approached only as $`r0`$. Thus the value of $`M_{matter}`$ calculated from here will be an upper bound. Applying this on the integral $`_{r_0}^{r_c}\beta \left(r\right)\rho _cr^2𝑑r`$ part Eq. 4.5 implies $`\left[\frac{1}{4}\left(r_c+r_0\right)\left(r_c^2+r_0^2\right)\frac{1}{3}r_0\left(\frac{r_c^3r_0^3}{r_cr_0}\right)\right]\frac{1}{4}r_c^3\rho _c`$ so that Eq. 4.5 reduces to $$_{matter}=\frac{4}{3}\pi \left[r_0^2r_g+\frac{1}{4}r_c^3\right]\left(\frac{\alpha }{3\left(\alpha +1\right)}\right)^{\frac{1}{2}}\rho _{\mathrm{max}}$$ (4.6) Use of $`r_c=\left[r_0^2r_g\mathrm{ln}\left(\frac{\rho _\epsilon }{\rho \mathrm{max}}\right)\right]^{\frac{1}{3}}`$ and $`\alpha =\mathrm{2.\hspace{0.17em}213\hspace{0.17em}5}`$ then gives $`\frac{_{2f}}{M}=\mathrm{0.479\hspace{0.17em}17}\left[1+\mathrm{0.735\hspace{0.17em}70}\right]=0.83160.83`$. This result agrees with the one found above in the quintessential picture. ## 5 Conclusion In this paper we have suggested a possible model for a non-singular black hole as a product of gravitational collapse. It is an exact solution of the Einstein Equations, being Type \[II, (II)\] by Petrov classification. This model is based on a judicious choice we have made of the equation of state of the collapsed matter. At high densities this equation of state violates some of the energy conditions originally used to justify the existence of black hole singularities. In our model this violation leads to non-singular collapse. On the other hand, for the entire density parameter space the equation of state in our approach still satisfies the Weak Energy Condition $`\rho 0`$, $`\rho +p_r0`$, a basic requirement for physical fields. The solution depicts a spacetime with matter fluids in the outer layers (region II) which, as one moves deeper inside, give way to either a quintessential or a two-fluid region III. Region III, in turn, evolves into region IV, an inner-most core with de Sitter characteristics. The fluids in regions III and IV both provide the negative pressures needed to sustain the outer matter in static equilibrium. The solution has the required asymptotic forms, reducing to the Schwarzschild vacuum solution outside the matter fields and reducing to the de Sitter solution as one approaches the black hole center. The existence of region III renders the interface between matter fields and the de Sitter vacuum to join smoothly. This is because both the density $`\rho `$, the radial pressure $`p_r`$ and the tangential pressure $`p_{}`$ profiles and their derivatives are continuous. This character of the fields makes the matching across associated interfaces natural. Using our model we have, in Section IV, offered two viable interpretations about the possible macro-state of the fields making up the total energy of the black hole. These interpretations are based on two pictures, namely a Quintessential Picture and a Two Fluid Picture. In both cases we have estimated the fractional contribution of the regular matter-like fields as a fraction of the total black hole energy. We find for an upper bound the same value of $`0.83`$in both cases<sup>3</sup><sup>3</sup>3It is remarkable that the two different interpretations yield the same numerical values of the mass-fractions.. As a corollary, our model suggests that the lower bound for the amount of matter that is found in the “exotic” state can be as small as $`17\%`$ of the entire configuration of the collapsed matter. Both the size and density of this de Sitter-like central core region will depend on the details of the microscopic physics that lead to the specific equation of state we have employed, which go beyond the scope of the present work. However, these considerations are consistent with our present notions of dynamical particle masses and symmetry breaking, as the density of this state is much higher than those produced to-date in the laboratory. The precise value is uncertain but it has to be at least as high as $`m_H^4`$, where $`m_H`$ is the mass of the Higgs field, and it can be as high as the Planck density. Our entire non-singular configuration is well within the horizon of the black hole and in this respect the configuration is very different from some previous treatments, e.g. . In these treatments, the field configuration fills the entire volume interior to $`r=2M`$ with a fluid with $`p=\rho `$, whose energy density for sufficiently large black holes can be smaller than that of water. This raises the difficult problem of converting matter from the usual equations of state to the “exotic” $`p=\rho ,`$ under usual laboratory conditions. Finally, the model leaves some issues unresolved. First, the calculation for these mass fractions is based on assuming that in the region $`r_0<r<R`$ both the pressure and the density are well-behaved, functions in the radial coordinate $`r`$, so that conversely $`r\left(\rho \right)=r_0^2r_g\mathrm{ln}\left(\frac{\rho }{\rho _{\mathrm{max}}}\right)`$ well defined in the entire parameter space of $`\rho `$. Since in our model $`\rho \rho _{\mathrm{max}}`$ when $`rr_0`$, then $`r\left(\rho \right)`$ vanishes at $`r_0`$ forcing (in the mass calculation) the de Sitter state to appear virtually at the origin. It is in this sense that $`\frac{_{matter}}{M}=0.83`$ is an upper-bound. One can reduce this fraction by using $`r\left(\rho \right)=r_0^2r_g\mathrm{ln}\left(\frac{\rho }{\rho _{\mathrm{max}}}\right)+r_0`$ in the calculations, instead. However since we don’t know the radial extent the de Sitter field could fill at the core, or whether indeed it fills any space bigger than the Planck length we can not presently constrain $`\frac{_{matter}}{M}`$from below. This is one issue for quantum gravity to settle. Secondly, our classical model can not distinguish between the two pictures offered to determine which choice of the matter macrostate is the correct one. However, the fact that in the region $`r_0<r<R`$ the energy-momentum tensor elements $`\rho `$ and $`p`$ are well-behaved functions of the radial coordinate $`r`$ may provide an interesting insight. It suggests that one can associate a unique value of the energy-momentum tensor on each 2-surface in the family $`\mathrm{\Sigma }=\left\{\mathrm{\Sigma }_{II}\mathrm{\Sigma }_{III}\right\}`$. One can further speculate that, such a configuration may even be quantizable. In such a case the information pertaining to the pre-collapse phase of the object would not be lost but would now reside on the (quantized) onion-like family of hyperfaces $`\mathrm{\Sigma }`$ inside the black hole. This is another issue for quantum gravity to settle. These results suggest a need for further investigation. ###### Acknowledgement 1 This research was performed while one of the authors (MRM) held a National Research Council Senior Research Associateship Award at (NASA-Goddard).
warning/0506/math0506116.html
ar5iv
text
# The problem of distinguishing between a center and a focus for nilpotent and degenerate analytic systems ## 1. Introduction and statement of the main results Two of the main and oldest problems in the qualitative theory of differential systems in $`^2`$ is the distinction between a center and a focus, called the center problem; and the determination of the first integrals in the case of centers, see for instance . This paper deals with these two problems for the class of analytic differential systems. Let $`p^2`$ be a singular point of a differential system in $`^2`$. We say that $`p`$ is a center if there is a neighborhood $`U`$ of $`p`$ such that all the orbits of $`U\{p\}`$ are periodic, and we say that $`p`$ is a focus if there is a neighborhood $`U`$ of $`p`$ such that all the orbits of $`U\{p\}`$ spiral either in forward or in backward time to $`p`$. Once we have a center at $`p`$ of a differential system in $`^2`$, another problem is to know if there exists or not a first integral $`H`$ defined in some neighborhood $`U`$ of $`p`$ (i.e. a non–constant function $`H:U`$ such that $`H`$ is constant on the orbits of the differential system), and to know the differentiability of $`H`$ with respect to the differentiability of the system. More specifically, we assume that we have an analytic differential system having a center at $`p`$. Then, it is known that there exists a $`C^{\mathrm{}}`$ first integral defined in some neighborhood of $`p`$, see . It is also known that there exists an analytic first integral defined in $`U\{p\}`$ for some neighborhood $`U`$ of $`p`$, see ; but such analytic first integral in general cannot be extended to $`p`$. For any center $`p`$ of an analytic differential system in $`^2`$ it is an open problem to characterize when there exists an analytic first integral in a neighborhood of $`p`$, or simply a local analytic first integral at $`p`$. A singular point $`p`$ is a monodromy singular point of a real analytic differential system in $`^2`$ if there is no characteristic orbit associated to it; i.e., there is no orbit tending to the singular point with definite tangent at this point. Let $`p`$ be a singular point of an analytic differential system. If $`p`$ is monodromy, then it is either a center or a focus, see . Moreover, $`p`$ is a center if and only if there exists a $`C^{\mathrm{}}`$ first integral defined in some neighborhood of $`p`$, see . Let $`p^2`$ be a singular point of an analytic differential system in $`^2`$, and assume that $`p`$ is a center. Without loss of generality we can assume that $`p`$ is the origin of coordinates (if necessary we do a translation of coordinates sending $`p`$ at the origin). Then, after a linear change of variables and a rescaling of the time variable (if necessary), the system can be written in one of the following three forms: (1) $`\dot{x}`$ $`=`$ $`y+F_1(x,y),\dot{y}=x+F_2(x,y);`$ (2) $`\dot{x}`$ $`=`$ $`y+F_1(x,y),\dot{y}=F_2(x,y);`$ (3) $`\dot{x}`$ $`=`$ $`F_1(x,y),\dot{y}=F_2(x,y);`$ where $`F_1(x,y)`$ and $`F_2(x,y)`$ are real analytic functions without constant and linear terms, defined in a neighborhood of the origin. In what follows a center of an analytic differential system in $`^2`$ is called linear type, nilpotent or degenerate if after an affine change of variables and a rescaling of the time it can be written as system (1), (2) or (3), respectively. The characterization of the linear type centers in terms of the existence of an analytic first integral is due to Poincaré and Liapunov , see also Moussu . Linear Type Center Theorem. The real analytic differential system (1) has a center at the origin if and only if there exists a local analytic first integral of the form $`H=x^2+y^2+F(x,y)`$ defined in a neighborhood of the origin, where $`F`$ starts with terms of order higher than $`2`$. An analytic system on the plane will have a singular point of center type if a countable number of conditions on the coefficients of the system are satisfied, see . Based on the Linear Type Center Theorem there is a method, called the Poincaré–Liapunov method, which consists in determining when a system of the form (1) has a local analytic first integral at the origin, and consequently a center at this point. This algorithm looks for a formal power series of the form (4) $$H(x,y)=\underset{n=2}{\overset{\mathrm{}}{}}H_n(x,y),$$ where $`H_2(x,y)=(x^2+y^2)/2`$, and for each $`n`$, $`H_n(x,y)`$ are homogeneous polynomials of degree $`n`$, so that (5) $$\dot{H}=\underset{k=2}{\overset{\mathrm{}}{}}V_{2k}(x^2+y^2)^k,$$ where the $`V_{2k}`$’s are called the Liapunov constants. It is known that the Liapunov constants are polynomials in the coefficients of system (1). We note that the Poincaré–Liapunov method for analytic differential systems is an algorithm which at each step uses only a finite jet of the system for the calculation of a Liapunov constant. The singular point is a center if and only if all the Liapunov constants vanish. For more details see and references therein. Until now there is no algorithm comparable to the Poincaré–Liapunov method for determining the center conditions in the case of nilpotent and degenerate singular points, except if the singular point has no characteristic direction because in this last case we can use the algorithm of Bautin (see also ). In any case the necessary computations for applying Bautin’s algorithm are in general more difficult to implement that the ones coming from the Poincaré–Liapunov method. In this paper we shall show that essentially the Poincaré–Liapunov algorithm also works for determining the analytic nilpotent centers and a subclass of the analytic degenerate centers. Our main result is the following one. ###### Theorem 1 (Nilpotent Center Theorem). Suppose that the origin of the real analytic differential system (2) is a center, then there exist analytic functions $`G_1`$ and $`G_2`$ without constants terms, such that the system (6) $$\dot{x}=y+F_1(x,y)+\epsilon xG_1(x,y),\dot{y}=\epsilon x+F_2(x,y)+\epsilon xG_2(x,y),$$ has a linear type center at the origin for all $`\epsilon >0`$. Roughly speaking Theorem 1 can be stated saying simply that an analytic nilpotent center is always limit of analytic linear type centers. Theorem 1 is proved in Section 2. By the Linear Type Center Theorem, system (6) has a local analytic first integral $`H_\epsilon (x,y)`$ at the origin for $`\epsilon >0`$. If there exists $$\underset{\epsilon 0}{lim}H_\epsilon (x,y),$$ and it is a function $`H(x,y)`$ well defined in a neighborhood of the origin, then $`H(x,y)`$ is a local first integral of system (2) at the origin. Note that, in general, $`H`$ is not analytic, see Remark 12. We note that the Nilpotent Center Theorem reduces the study of the nilpotent centers to the case of linear type centers. So, we can apply the Poincaré–Liapunov method to system (6), looking for analytic first integrals of the form $`H=(\epsilon x^2+y^2)/2+F(x,y,\epsilon )`$, where $`F`$ starts with terms of order higher than $`2`$ in the variables $`x`$ and $`y`$. We determine the Liapunov constants $`V_{2k}`$ from (5). Several examples showing the application of the Poincaré–Liapunov method to detect nilpotent centers are given in Section 4. Based in the results obtained for nilpotent centers we establish the following definition. Suppose that the origin of the real analytic differential system (3) is a center. We say that it is limit of linear type centers if there exist $`G_1`$ and $`G_2`$ analytic functions in $`x`$, $`y`$ and $`\epsilon `$, without constants and linear terms in $`x`$ and $`y`$, such that the system (7) $$\dot{x}=\epsilon y+F_1(x,y)+\epsilon G_1(x,y,\epsilon ),\dot{y}=\epsilon x+F_2(x,y)+\epsilon G_2(x,y,\epsilon ),$$ has a linear type center at the origin for all $`\epsilon 0`$ sufficiently small. A more general definition of limit of linear type centers would be to consider functions $`G_1`$ and $`G_2`$ that are not analytic in $`\epsilon `$. ###### Theorem 2. Suppose that the origin of the real analytic differential system (2) or (3) is monodromy, and that this system is limit of linear type centers of the form (6) or (7), respectively. Suppose also that there are no singular point of (6) or (7) tending to the origin when $`\epsilon `$ tends to zero. Then, system (2) or (3) has a center at the origin. Theorem 2 is proved in Section 2. The condition that there are no singular point tending to the origin when $`\epsilon `$ tends to zero is easily verifiable using the lower order terms of the perturbed system (6) or (7). Another difficulty of the problem of distinguishing between a center and a focus becomes from the fact that this problem for degenerate centers can be no algebraically solvable; i.e., it does not exist an infinite sequence of independent polynomial expressions involving the coefficients of the system, such that their simultaneous vanishing guarantees the existence of a center, see . The problem of distinguishing between a center and a focus is algebraically solvable in the class of analytic differential systems of type (1) and (2), see . The Nilpotent Center Theorem provides a new proof of the fact that nilpotent analytic centers are algebraically solvable. This is due to the fact that we have seen in Theorem 1 that the nilpotent centers can be $`\epsilon `$–approximated by systems having linear type centers. Then, applying the Poincaré–Liapunov method to these linear type centers, and doing the limit when $`\epsilon 0`$ we obtain algebraic conditions characterizing the existence of a nilpotent center. So, in fact Theorem 1 provides an algorithm for solving the center problem for nilpotent centers. See the end of Section 2 for more details, and also Section 4. For centers of the form (3) some preliminary results exist for distinguishing between a center and a focus, see for instance . We say that an analytic differential system in the plane is time–reversible (with respect to an axis of symmetry through the origin) if after a rotation $$\begin{array}{cccc}\hfill \left(\begin{array}{c}\xi \\ \eta \end{array}\right)& =& \left(\begin{array}{cc}\mathrm{cos}\alpha & \mathrm{sin}\alpha \\ \mathrm{sin}\alpha & \mathrm{cos}\alpha \end{array}\right)& \left(\begin{array}{c}x\\ y\end{array}\right),\hfill \end{array}$$ the system in the new variables $`(\xi ,\eta )`$ becomes invariant by a transformation of the form $`(\xi ,\eta ,t)(\xi ,\eta ,t)`$. The phase portrait of this new system is symmetric with respect to the straight line $`\xi =0`$. We note that for all reversible nilpotent centers which are symmetric with respect to a straight line through the origin, this line can be only the line of the axes $`x`$ or $`y`$. We remark that all the nilpotent centers that we know are time–reversible or have an analytic first integral at the origin. In the case of degenerate centers is much more difficult to distinguish between a center and a focus than in the case of linear and nilpotent type centers. In the next theorem we present some results for the degenerate centers. ###### Theorem 3. For a degenerate analytic center the following statements hold. * A Hamiltonian degenerate center is always limit of linear type Hamiltonian centers. * A time–reversible degenerate center is always limit of linear type time–reversible centers. * There are degenerate centers which are neither Hamiltonian nor time–reversible that are limit of linear type centers. * Non algebraically solvable degenerate centers are not limit of linear type centers. * There are algebraically solvable degenerate centers which are not limit of linear type centers. * There exist degenerate centers with characteristic directions which are limit of degenerate centers without characteristic directions. Theorem 3 is proved in Section 3. Let (3) be a family of analytic systems depending on several parameters. Inside the degenerate centers of this family we can determine those which are limit of linear type centers of the form (7). For this kind of systems we can apply the Poincaré–Liapunov method to system (7) with $`\epsilon 0`$ and compute their Liapunov constants. Vanishing these Liapunov constants we obtain the center conditions for the system (7). Taking the limit when $`\epsilon 0`$ in these conditions, we get the center conditions for the degenerate centers (3). System (7) must be a linear type center only for $`\epsilon 0`$ sufficiently small. In consequence, in the applications of the Poincaré–Liapunov algorithm, it is sufficient to calculate the Liapunov constants up to first order in $`\epsilon `$. In contrast, for the nilpotent centers, which are always limit of linear type centers, we can calculate up to any order in $`\epsilon `$ because system (6) has a center at the origin for all $`\epsilon >0`$, and from this fact we can obtain several conditions at each step of the algorithm. In Section 4 we provide an example of the application of this method to a family of polynomial differential systems (3). Finally, in Section 5 we obtain some results on the cyclicity of nilpotent and degenerate centers which are limit of linear type centers. In particular, we prove the following result (for a definition of cyclicity of a center see Section 5). ###### Proposition 4. Consider a nilpotent center or a degenerate center of a polynomial differential system (2) or (3) of degree $`m`$. We suppose that this center is limit when $`\epsilon 0`$ of linear type centers of polynomial differential systems of degree $`n`$ of the form (6) or (7), respectively. If the Liapunov constants of a general perturbation of the same degree $`n`$ of the linear type centers (6) or (7) are well–defined when $`\epsilon 0`$ and the Poincaré map for a perturbation of the initial nilpotent or degenerate center is analytic, then the following statements hold: * The cyclicity of the nilpotent center (2) is at most the cyclicity of the linear type center (6) for all $`\epsilon >0`$. * The cyclicity of the degenerate center (3) is at most the cyclicity of the linear type centers (7) for $`\epsilon 0`$ sufficiently small. ## 2. Proof of Theorems 1 and 2 The characterization of the nilpotent centers in terms of the existence of a symmetry is due to Berthier and Moussu who obtained the following result. We shall need it in the proof of Theorem 1. ###### Theorem 5. If the analytic system (2) has a center at the origin, then there exists an analytic change of variables such that the new system has also the form (2) and it is invariant by the change of variables $`(x,y,t)(x,y,t)`$. We recall from that if the analytic system (2) has a center at the origin and there exists an analytic change of variables such that the new system has also the form (2) and it is invariant by the change of variables $`(x,y,t)(x,y,t)`$, then the system has an analytic first integral defined in a neighborhood of the origin. Proof of Theorem 1: Assume that the origin of system (2) is a center. Theorem 5 and its proof says that for any nilpotent center (2) corresponding to an analytic vector field $`X(x,y)`$, there exists an analytic change of variables $`(x,y)(u,v)`$ of the form (8) $$x=u+\mathrm{},y=v+\mathrm{},$$ such that $`X(x,y)`$ written in the new variables is a vector field of the form (9) $$Y(u,v)=(v+\overline{F}_1(u,v),\overline{F}_2(u,v)),$$ where $`\overline{F}_1`$ and $`\overline{F}_2`$ are analytic functions starting with terms of second degree in $`x`$ and $`y`$, and the associated differential system is invariant under the change of variables $`(u,v,t)(u,v,t)`$. Now we consider the following perturbation of the vector field (9): (10) $$Y_\epsilon (u,v)=(v+\overline{F}_1(u,v),\epsilon u+\overline{F}_2(u,v)),$$ with $`\epsilon >0`$. Since the eigenvalues at the singular point located at the origin are $`\pm \sqrt{\epsilon }i`$, and the differential system associated to the vector field (10) is invariant under the change of variables $`(u,v,t)(u,v,t)`$ (because the unperturbed system is invariant), it follows that the origin of the vector field (10) is a linear type center for all $`\epsilon >0`$. Using the inverse of the change of variables (8) we get that the differential system associated to the vector field (10) becomes (11) $$\dot{x}=y+F_1(x,y)+\epsilon xG_1(x,y),\dot{y}=\epsilon x+F_2(x,y)+\epsilon xG_2(x,y),$$ where $`G_1`$ and $`G_2`$ are analytic functions without constants terms, depending on the change of variable (8). Let $`X_\epsilon (x,y)`$ be the vector field associated to system (11). Since $`Y_\epsilon (u,v)`$ has a linear type center at the origin for all $`\epsilon >0`$, the same holds for $`X_\epsilon (x,y)`$. This completes the proof of the theorem. Proof of Theorem 2: Consider an analytic system $`(P,Q)`$ of the form (2) or (3) with a monodromy singular point $`p`$ at the origin. Suppose that this system is limit of linear type centers $`(P_\epsilon ,Q_\epsilon )`$ of the form (6) or (7), respectively. Since the origin is monodromy, if $`S`$ is a sufficiently small curve with an endpoint at the origin, then the Poincaré map $`\mathrm{\Pi }:SS`$ associated to the system $`(P,Q)`$ is well–defined and the leading term is always linear for a suitable choice of a semi–transversal algebraic curve, which can have a singularity at the singular point, see . The Poincaré map $`\mathrm{\Pi }_\epsilon :SS`$ associated to the system $`(P_\epsilon ,Q_\epsilon )`$ is the identity for all $`\epsilon >0`$ if the center is nilpotent, and for $`\epsilon `$ sufficiently small and $`\epsilon 0`$ if the center is degenerate. Therefore, by the theorem on analytic dependence on initial conditions and parameters, it follows that $`\mathrm{\Pi }=lim_{\epsilon 0}\mathrm{\Pi }_\epsilon `$ if the center is nilpotent, or $`\mathrm{\Pi }=lim_{\epsilon 0}\mathrm{\Pi }_\epsilon `$ if the center is degenerate. Hence, we conclude that $`\mathrm{\Pi }`$ is the identity. So, the monodromy singular point $`p`$ of $`(P,Q)`$ is a center. The condition that there are not singular points tending to the origin when $`\epsilon `$ tends to zero guarantees that the domain of $`\mathrm{\Pi }_\epsilon `$ does not reduce to the origin when $`\epsilon `$ tends to zero. Theorems 1 and 2 can be used to detect nilpotent centers of analytic differential systems applying the algorithm of Poincaré–Liapunov. In the particular case of polynomial systems the method works as follows. We consider the system (12) $$\dot{x}=y+F_1(x,y),\dot{y}=F_2(x,y),$$ where $`F_1`$ and $`F_2`$ are polynomials without constants and linear terms containing a set of arbitrary parameters and such that the origin is a monodromy singular point. We recall that using Andreev’s Theorem we can know when a nilpotent singular point is or not monodromy, see . For detecting the centers of (12), according with Theorem 1, we consider the perturbed system (13) $$\dot{x}=y+F_1(x,y)+\epsilon xG_1(x,y),\dot{y}=\epsilon x+F_2(x,y)+\epsilon xG_2(x,y),$$ where $`xG_1`$ and $`xG_2`$ are analytic functions starting with quadratic terms in $`x`$ and $`y`$. We apply now the Poincaré–Liapunov algorithm to determine necessary conditions to have a center at the origin for system (13). In general, these conditions will be satisfied by choosing conveniently the coefficients of the analytic functions $`G_1`$ and $`G_2`$. When this is not possible we must employ the parameters of the polynomial system (12). In this way we will obtain necessary conditions for the existence of a center at the origin of system (12). The set of sufficient conditions of center for the non–perturbed system (12) will be obtained in a finite number of steps, because the Hilbert’s basis theorem guarantees that this process is finite. Every time that we find a necessary condition for the non–perturbed system (12) we must to study if the non-perturbed system (12) already have a center at the origin. As the number of steps is finite and for determining each Poincaré–Liapunov constant of the perturbed system (13) we need only a finite jet, the necessary perturbation to detect the center cases will be polynomial, i.e., the functions $`G_1`$ and $`G_2`$ will be polynomials. We note that, under the assumptions of Theorem 2, it is not possible to satisfy the center conditions of (13) only with the parameters of the perturbation because in that case using Theorem 2 the nilpotent polynomial system would have a center for arbitrary values of the parameters of the family, which is a contradiction if the initial system (12) has not a center at the origin. ## 3. Proof of Theorem 3 In this section we shall work with an analytic degenerate center (3) defined in a neighborhood of the origin. Proof of Theorem 3(a): Suppose that system (3) is Hamiltonian with Hamiltonian $`H=H(x,y)`$. The system (14) $$\dot{x}=\epsilon y+F_1(x,y),\dot{y}=\epsilon x+F_2(x,y),$$ is also a Hamiltonian system with the Hamiltonian first integral $`\epsilon (x^2+y^2)/2+H(x,y)`$. Consequently, system (14) has a linear type center at the origin for $`\epsilon 0`$, and the initial degenerate center (3) is obtained taking in system (14) the limit when $`\epsilon 0`$ . Proof of Theorem 3(b): Without loss of generality, taking into account the definition of a time–reversible system, we can assume that system (3) is invariant by the change of variables $`(x,y,t)(x,y,t)`$. Consider the perturbation of it given by a system of the form (14). Then, it is easy to see that system (14) is also invariant under the change of variables $`(x,y,t)(x,y,t)`$. Therefore, since the eigenvalues of the linear part at the origin of system (14) are $`\pm \sqrt{|\epsilon |}i`$, it has a linear type center at the origin for $`\epsilon 0`$. Again, the initial degenerate center (3) is obtained taking in system (14) the limit when $`\epsilon 0`$ . Proof of Theorem 3(c): Consider the following quartic polynomial differential system (15) $$\begin{array}{ccc}\dot{x}\hfill & =\hfill & (y+y^2)(x^2+y^2),\hfill \\ \dot{y}\hfill & =\hfill & (x+2x^2)(x^2+y^2).\hfill \end{array}$$ It is easy to see that this system has a degenerate center at the origin, because removing the common factor $`x^2+y^2`$ (doing a change of the independent variable) we get a quadratic Hamiltonian system having a center at the origin. It is easy to check that system (15) is not Hamiltonian, and that it has the first integral (16) $$H(x,y)=(x^2+y^2)/2+2x^3/3y^3/3.$$ Now we claim that system (15) is not time–reversible. Suppose that it is time–reversible. Then, there exists a rotation which pass the variables $`(x,y)`$ to the new variables $`(u,v)`$, given by (17) $$u=\mathrm{cos}\alpha x\mathrm{sin}\alpha y,v=\mathrm{sin}\alpha x+\mathrm{cos}\alpha y,$$ which transforms the axis of symmetry into the line $`u=0`$. In the new variables the system becomes $`\dot{u}=P(u,v)`$ and $`\dot{v}=Q(u,v)`$. Since this system must be invariant by $`(u,v,t)(u,v,t)`$, we must have $$P(u,v)=P(u,v),Q(u,v)=Q(u,v).$$ These two equations are satisfied if and only if $$\mathrm{cos}\alpha \mathrm{sin}\alpha (2\mathrm{cos}\alpha \mathrm{sin}\alpha )=0,2\mathrm{sin}^3\alpha \mathrm{cos}^3\alpha =0.$$ Since this system has no solution, the claim is proved. Now, consider the following perturbation of system (15) (18) $$\begin{array}{ccc}\dot{x}\hfill & =\hfill & (y+y^2)(x^2+y^2\epsilon ),\hfill \\ \dot{y}\hfill & =\hfill & (x+2x^2)(x^2+y^2\epsilon ).\hfill \end{array}$$ System (18) has also the first integral (16). Consequently, system (18) has a center at the origin for all $`\epsilon `$. This center is of linear type if $`\epsilon 0`$. Hence, doing the limit $`\epsilon 0`$ in system (18), we obtain the initial system (15) with a degenerate center at the origin. We have seen in Theorem 1 that the nilpotent centers can be $`\epsilon `$–approximated by systems having linear type centers. Then, applying the Poincaré–Liapunov method to these linear type centers, and doing the limit when $`\epsilon 0`$ we obtain algebraic conditions characterizing the existence of a nilpotent center. We shall see this more explicitly in Section 4. Now, we shall show that this does not occur for degenerate centers which are not algebraically solvable (proving statement (d) of Theorem 3), and for some classes of degenerate centers which are algebraically solvable (proving statement (e) of Theorem 3). Proof of Theorem 3(d): It is known that the problem of determining the center conditions at the origin of the system $`\dot{x}`$ $`=`$ $`xp_2yp_1+4x(x^2+\mu y^2)p_1,`$ $`\dot{y}`$ $`=`$ $`xp_1+yp_2+4y(x^2+\mu y^2)p_1,`$ where $`p_1=x^2+a_4xy+a_5y^2`$ and $`p_2=a_1x^2+a_2xy+a_3y^2`$, is not algebraically solvable for some specific values of the parameters, see . If such a system is limit of linear type centers, it would be algebraically solvable. So, the proof of statement (d) of Theorem 3 follows. Proof of Theorem 3(e): Consider the following cubic homogeneous system (19) $$\begin{array}{ccc}\dot{x}\hfill & =\hfill & P(x,y)=12\lambda x^39x^2y20\lambda xy^225y^3+9\mu y^3,\hfill \\ \dot{y}\hfill & =\hfill & Q(x,y)=9x^3+12\lambda x^2y+25xy^220\lambda y^3,\hfill \end{array}$$ with the monodromy condition that $`xQ(x,y)`$ $`yP(x,y)`$ has no real factors. This system has a degenerate center at the origin if and only if $`\mu =0`$, or $`\lambda =0`$. This follows checking the conditions (i) and (ii) of the Appendix which characterize the homogeneous systems having a center at the origin. Condition (i) is satisfied by the monodromy condition, and condition (ii) is $`{\displaystyle _0^{2\pi }}{\displaystyle \frac{f(\theta )}{g(\theta )}}𝑑\theta =0`$; i.e. $`{\displaystyle \frac{4\pi \lambda }{\sqrt{9\mu 25}\sqrt{81\mu +64}\sqrt{17\sqrt{64+81\mu }}\sqrt{17+\sqrt{64+81\mu }}}}`$ $`(\sqrt{17\sqrt{64+81\mu }}(16027\mu 5\sqrt{64+81\mu })+`$ $`\sqrt{17+\sqrt{64+81\mu }}(160+27\mu 5\sqrt{64+81\mu }))=0.`$ It is easy to check that condition (ii) holds if and only if $`\mu =0`$ or $`\lambda =0`$. Moreover, these centers are algebraically solvable because condition (ii) is algebraic. In the case $`\mu =0`$ condition (i) is directly satisfied because $`xQ(x,y)yP(x,y)=(x^2+y^2)(9x^2+25y^2)`$. In the case $`\lambda =0`$, the homogeneous polynomial $`xQ(x,y)yP(x,y)`$ is $`9x^4+34x^2y^2+25y^49\mu y^4`$, and condition (i) is satisfied if and only if $`\mu <25/9`$. We consider the following perturbation of system (19) (20) $$\begin{array}{cc}& \dot{x}=\epsilon y+12\lambda x^39x^2y20\lambda xy^225y^3+9\mu y^3+\epsilon G_1,\hfill \\ & \dot{y}=\epsilon x+9x^3+12\lambda x^2y+25xy^220\lambda y^3+\epsilon G_2,\hfill \end{array}$$ where $`G_i=G_i(x,y,\epsilon )`$, for $`i=1,2`$, are analytic functions in $`x`$, $`y`$ and $`\epsilon `$, without constants and linear terms in $`x`$ and $`y`$. Applying the Poincaré–Liapunov method to system (LABEL:ss6) (see Section 4 for more details), we obtain that the first Liapunov constant is $$V_1=8\lambda +\epsilon \overline{V}_1,$$ where $`\overline{V}_1`$ is the first Liapunov constant of the analytic system $$\begin{array}{ccc}\dot{x}\hfill & =\hfill & \epsilon y+\epsilon G_1(x,y,\epsilon ),\hfill \\ \dot{y}\hfill & =\hfill & \epsilon x+\epsilon G_2(x,y,\epsilon ).\hfill \end{array}$$ As we must vanish $`V_1`$ up to first order in $`\epsilon `$ we obtain only the condition $`\lambda =0`$. So, the case $`\mu =0`$ and $`\lambda `$ cannot be detected as limit of linear type centers. Proof of Theorem 3(f): We consider the system (21) $$\dot{x}=ay^3,\dot{y}=bx^5,$$ with $`ab>0`$. It is a Hamiltonian system with Hamiltonian $$H(x,y)=\frac{ay^4}{4}+\frac{bx^6}{6}.$$ It is easy to check that system (21) is a $`(2,3)`$–quasi–homogeneous system of weight degree $`8`$ satisfying conditions (i) and (ii) for having a degenerate center at the origin, see the Appendix. Another way to see that the origin is a degenerate center is noting that the level curves of $`H`$ are ovals. It is easy to check that system (21) has only one characteristic direction, given by $`y=0`$. Now, consider the following perturbation of system (21): (22) $$\dot{x}=ay^3,\dot{y}=\epsilon x^3+bx^5,$$ with $`a\epsilon >0`$. This system is also Hamiltonian, with $$H_\epsilon (x,y)=\frac{ay^4}{4}+\frac{\epsilon x^4}{4}+\frac{bx^6}{6}.$$ System (22) has a degenerate center at the origin, because the origin is surrounded by ovals. This system has no characteristic direction. Now, doing the limit $`\epsilon 0`$ in system (22), we obtain the initial system (21) with a degenerate center at the origin and with a characteristic direction. The example given in the proof of Theorem 3(f) shows that, in a similar way that we can apply the Poincaré–Liapunov method to detect nilpotent centers, in the study of certain degenerate centers with characteristic directions we can apply the Bautin method for degenerate centers without characteristic directions (see for instance ) to a convenient perturbation of the system with characteristic directions. ## 4. The Poincaré–Liapunov method for nilpotent and degenerate systems In this section we illustrate how to apply the Poincaré–Liapunov method to several families of polynomial differential systems for detecting nilpotent or degenerate centers. Some of these families have been studied recently by other authors with different and more complicated techniques. First we start studying some nilpotent centers. We note that the simplest nilpotent polynomial centers must be of degree $`3`$, because there are no nilpotent center for quadratic polynomial differential systems, see for instance . We consider the system (23) $$\dot{x}=y+x^2+k_2xy,\dot{y}=k_1x^2x^3.$$ We apply to this family the general algorithm, with a general perturbation and we obtain the following result: ###### Proposition 6. System (23) has a nilpotent center at the origin if and only if $`k_1=k_2=0`$. Proof: Applying the Poincaré–Liapunov method to the perturbed system (24) $$\dot{x}=y+x^2+k_2xy+\epsilon xG_1(x,y),\dot{y}=\epsilon x+k_1x^2x^3+\epsilon xG_2(x,y),$$ where $$G_1(x,y)=\underset{i+j1}{\overset{\mathrm{}}{}}a_{ij}x^iy^j,G_2(x,y)=\underset{i+j1}{\overset{\mathrm{}}{}}b_{ij}x^iy^j,$$ and $`\epsilon >0`$, we obtain the first Liapunov constant $$V_1=\frac{2}{3+2\epsilon +3\epsilon ^2}[2k_1+(2b_{10}+2a_{10}k_1+b_{01}k_1k_2)\epsilon $$ $$(a_{01}3a_{20}2a_{10}b_{10}b_{01}b_{10}b_{11}+a_{10}k_2)\epsilon ^2+(a_{02}a_{01}a_{10})\epsilon ^3].$$ We note that in $`V_1`$ only appear the linear and quadratic terms of $`G_1`$ and $`G_2`$. Vanishing $`V_1`$ at any order in $`\epsilon `$ we get the necessary center condition $`k_1=0`$ for system (23) and for an arbitrary perturbation. We obtain also the conditions $`b_{10}=k_2/2`$, $`a_{01}=(6a_{20}+2b_{11}+b_{01}k_2)/2`$, and $`a_{02}=(6a_{10}a_{20}2a_{10}b_{11}+a_{10}b_{01}k_2)/2`$ on the parameters of the perturbation. The next Liapunov constant has the form $$V_2=\frac{1}{(1+\epsilon )(52\epsilon +5\epsilon ^2)}\left[72k_2+O(\epsilon )\right].$$ In the expression of $`V_2`$ we have contributions of the linear, quadratic, cubic, quartic and quintic terms of $`G_1`$ and $`G_2`$. Therefore, the conditions $`k_1=k_2=0`$ are necessary in order that the origin of system (23) be a center. These conditions are also sufficient as it is explained in Remark 12. We see that it has been sufficient to employ a polynomial perturbation of degree $`5`$ ir order to determine the necessary and sufficient conditions of center for system (23). Although in Theorem 1 the perturbation is unknown, it is surprising that with the simple perturbation $`\epsilon x`$ in $`\dot{y}`$ it is possible to obtain the center cases of many families, as it will be shown in the following examples. We consider the system (25) $$\dot{x}=y+Axy+By^2,\dot{y}=x^3+Kxy+Ly^3.$$ Applying the Andreev results we can see that the origin of system (25) is monodromy. ###### Proposition 7. System (25) has a nilpotent center at the origin if and only if $`AB3L=0`$ and $`AB(A^22K)=0`$. Proof: Applying the Poincaré–Liapunov method to the perturbed system (26) $$\dot{x}=y+Axy+By^2,\dot{y}=\epsilon xx^3+Kxy+Ly^3,$$ with $`\epsilon >0`$, we obtain the first Liapunov constant $$V_1=\frac{2\epsilon ^2(AB3L)}{3+2\epsilon +3\epsilon ^2}.$$ Vanishing $`V_1`$ we get the first center condition $`L=AB/3`$. Now, we compute the second Liapunov constant $$V_2=\frac{2\epsilon ^2AB(A^22K)}{3(1+\epsilon )(52\epsilon +5\epsilon ^2)}.$$ Vanishing $`V_2`$ we obtain the second center condition $`AB(A^22K)=0`$. So, these two conditions are necessary in order that the origin of the perturbed system (26) be a center. These two conditions are not necessary, in principle, for system (25), because we must investigate for others polynomials perturbations of the form $$\dot{x}=y+Axy+By^2+\epsilon xG_1(x,y),\dot{y}=\epsilon xx^3+Kxy+Ly^3+\epsilon xG_2(x,y).$$ But, in it is proved that these two conditions are necessary in order that the origin of system (25) be a center. We remark that in this particular system we do not need to take $`\epsilon =0`$ in the center conditions because they are independent of $`\epsilon `$. Now we prove that these two conditions are sufficient. If $`A=0`$ or $`B=0`$ (and consequently $`L=0`$), we have that system (25) is reversible with respect to $`(x,y,t)(x,y,t)`$ or $`(x,y,t)(x,y,t)`$, respectively. Therefore, since the origin is monodromy, it is a center. If $`AB0`$, $`L=AB/3`$ and $`A^22K=0`$, then the system has the analytic first integral $$H=\mathrm{exp}(Ax)\left(y^2\frac{12}{A^4}\frac{12}{A^3}x\frac{6}{A^2}x^2\frac{2}{A}x^3+Axy^2+\frac{2}{3}By^3\right).$$ This first integral can be obtained using the theory of integrability of Darboux, see for instance . In fact, this first integral already appeared in . Since the origin is monodromy, by the existence of this analytic first integral defined at the origin it follows that the origin is a center. We note that in this case the nilpotent center is neither time–reversible nor Hamiltonian. We consider the system (27) $$\dot{x}=y,\dot{y}=x^5+ax^6+y(bx^3+cx^4).$$ Applying Andreev’s results we can see that the origin of system (27) is monodromy. ###### Proposition 8. System (27) has a nilpotent center at the origin if and only if $`ab=0`$ and $`c=0`$. Proof: Applying the Poincaré–Liapunov method to the perturbed system (28) $$\dot{x}=y,\dot{y}=\epsilon x+x^5+ax^6+y(bx^3+cx^4),$$ with $`\epsilon >0`$, we obtain the first Liapunov constant $$V_1=\frac{2\epsilon c}{5+3\epsilon +3\epsilon ^2+5\epsilon ^3}.$$ Vanishing $`V_1`$ we get the first center condition $`c=0`$. Now, we compute the second Liapunov constant $$V_2=\frac{(2+7\epsilon )ab}{128\epsilon ^2}.$$ Vanishing $`V_2`$ we obtain the second center condition $`ab=0`$. So, these two conditions are necessary in order that the origin of the perturbed system (28) be a center. These two conditions are not necessary, in principle, for system (27) because we must investigate, as in the previous example, for others polynomials perturbations of the form $$\dot{x}=y+\epsilon xG_1(x,y),\dot{y}=\epsilon x+x^5+ax^6+y(bx^3+cx^4)+\epsilon xG_2(x,y).$$ But, in it is proved that these two conditions are necessary in order that the origin of system (27) be a center. Now we prove that these two conditions are sufficient. If $`a=c=0`$ or $`b=c=0`$, we have that the system is reversible with respect to $`(x,y,t)(x,y,t)`$ or $`(x,y,t)(x,y,t)`$, respectively. Therefore, since the origin is monodromy, it is a center. Proposition 7 is proved in by using Liapunov polar coordinates, see , and computing some generalized Liapunov constants. The method developed in is not useful to solve the center problem of Proposition 8, see . Proposition 8 is proved in by using the normal form theory and taking into account that a convenient truncated normal form of the nilpotent system is a Lienard system. The method developed in this paper solves both problems in a unified form and in a more simple way, by computing the Poincaré–Liapunov constants of a linear center type system. In both proofs we have used the results of and to prove that the conditions are necessary. This is not a restriction of our method because we can apply it with a general perturbation. But, in that case it is necessary to make a big amount of computations for obtaining the necessary conditions. This is the usual amount of computations that appear in the application of the Poincaré–Liapunov method when the system under study has several parameters. We consider the system (29) $$\dot{x}=y+a_{11}xy+a_{02}y^2+a_{30}x^3+a_{21}x^2y+a_{12}xy^2+a_{03}y^3,\dot{y}=x^3.$$ ###### Proposition 9. System (29) has a nilpotent center at the origin if and only if $`a_{30}=0`$, $`a_{02}a_{11}+a_{12}=0`$, $`a_{02}a_{11}a_{21}=0`$, and $`a_{02}a_{11}a_{03}=0`$. We consider the system (30) $$\dot{x}=y,\dot{y}=a_{11}xy+a_{02}y^2+a_{30}x^3+a_{21}x^2y+a_{12}xy^2+a_{03}y^3.$$ ###### Proposition 10. System (30) has a nilpotent center at the origin if and only if $`a_{21}a_{02}a_{11}=0`$, $`a_{03}=0`$, $`a_{02}a_{11}a_{30}=0`$, and $`a_{02}a_{11}(3a_{02}^2+2a_{12})=0`$. The proofs of Propositions 9 and 10 are similar to the proof of Propositions 7 and 8 and we omit them. Proposition 9 is also proved in and Proposition 10 is proved in . As the previous examples show, in some cases, it is sufficient to perturb system (2) with $`\epsilon x`$ in $`\dot{y}`$, but there are nilpotent centers which are limit of more general perturbations and which cannot be detected only with the perturbation $`\epsilon x`$ in $`\dot{y}`$, as we will see in the following example. We consider the system (31) $$\begin{array}{cc}& \dot{x}=P(x,y)=y+xy+(1a)y^2+(1a)xy^2ax^4ax^5,\hfill \\ & \dot{y}=Q(x,y)=cy^22x^3+cy^32x^3y+(c2)x^4(1+y).\hfill \end{array}$$ Applying the Andreev’s Theorem it is easy to see that system (31) has a monodromy singular point at the origin. ###### Proposition 11. System (31) has a nilpotent center at the origin for all values of $`a`$ and $`c`$. Proof: System (31) has the following analytic first integral $$H(x,y)=(1+x)^{2c}(1+y)^{2a}(x^4+y^2),$$ which can be determined using the Darboux theory of integrability. Therefore, system (31) has a center at the origin. Applying the Poincaré–Liapunov method to the perturbed system (32) $$\begin{array}{ccc}\dot{x}\hfill & =\hfill & P(x,y),\hfill \\ \dot{y}\hfill & =\hfill & \epsilon x+Q(x,y).\hfill \end{array}$$ with $`\epsilon >0`$, we obtain the first Liapunov constant $$V_1=\frac{2\epsilon ^2c(1+2a)}{3+2\epsilon +3\epsilon ^2}.$$ Therefore, the first center condition is $`c(1+2a)=0`$. But, we know that system (31) has a nilpotent center for all values of $`a`$ and $`c`$. Then, it must exist another more general $`\epsilon `$–perturbation of system (31) which is a linear type center for all values of $`a`$ and $`c`$. Consider the following polynomial perturbed system (33) $$\begin{array}{ccc}\dot{x}\hfill & =\hfill & \epsilon x(ax+ax^2)+P(x,y),\hfill \\ \dot{y}\hfill & =\hfill & \epsilon x(1+(1c)x+y+(1c)xy)+Q(x,y),\hfill \end{array}$$ where $`P`$ and $`Q`$ are defined in system (31). System (33) has a linear type center at the origin because it has the following analytic first integral $$H(x,y)=(1+x)^{2c}(1+y)^{2a}(x^4+y^2+\epsilon x^2).$$ Therefore, all the nilpotent centers of system (31) are limit of the linear type centers of system (33), but not all the nilpotent centers of system (31) are limit of linear type centers of system (32). This example shows that it is not always possible to obtain a nilpotent center as a limit of linear type centers with the only perturbation $`\epsilon x`$ in $`\dot{y}`$, even in the case where a local analytic first integral exists. ###### Remark 12. Consider the system (34) $$\dot{x}=y+x^2,\dot{y}=x^3.$$ Since this system is time–reversible with respect to the change of variables $`(x,y,t)(x,y,t)`$, and the origin is monodromy (see ), it has a nilpotent center at the origin. But it has neither a local analytic first integral, nor a formal first integral defined at the origin, see the proof in . Consider now the following perturbation of system (34). (35) $$\dot{x}=y+x^2,\dot{y}=\epsilon xx^3.$$ As this system is time–reversible with respect to the same change of variables, for $`\epsilon >0`$ it has also a center at the origin. Therefore, by the Linear Type Center Theorem we know that system (35) has a local analytic first integral $`H(x,y,\epsilon )`$ at the origin. It is possible to compute an explicit expression of it given by $$\mathrm{exp}\left[2\mathrm{arg}(\epsilon +x^2+i(x^2+2y\epsilon ))\right](\epsilon ^2+x^42\epsilon y+2x^2y+2y^2).$$ Now, taking the limit when $`\epsilon 0`$, we obtain the first integral $$H(x,y)=\underset{\epsilon 0}{lim}H(x,y,\epsilon )=\mathrm{exp}\left[2\mathrm{arg}(x^2+i(x^2+2y))\right](x^4+2x^2y+2y^2),$$ of system (34), which is not analytic at the origin. ∎ We see in this example that the limit of an analytic first integral defined in a neighborhood of the origin can be not analytic. In general the study of the nilpotent centers is easier with the algorithm proposed in this work than applying the results of . In our case, we have two arbitrary functions $`G_1`$ and $`G_2`$, while in the algorithm consequence of the results of there are three arbitrary functions, the one which appear in the normal form for the nilpotent center and the two coming from the change of variables. Moreover, for polynomial systems, under the assumptions of Theorem 2, the two arbitrary functions $`G_1`$ and $`G_2`$ of our method are always polynomials and this fact does not happen in the algorithm based on the results of . Now we apply the Poincaré–Liapunov method to detect degenerate centers in a family of polynomial differential systems. We consider the polynomial system (36) $$\begin{array}{ccc}\dot{x}\hfill & =\hfill & a(1+x)(x^44y^33y^4)+\mu y^3,\hfill \\ \dot{y}\hfill & =\hfill & a(1+y)(4x^3+3x^4y^4)+\lambda x^5.\hfill \end{array}$$ with the monodromy condition $`a\mu >0`$ if $`a0`$ and $`\mu \lambda <0`$ if $`a=0`$. ###### Proposition 13. System (36), with the monodromy condition $`a\mu >0`$ if $`a0`$ and $`\mu \lambda <0`$ if $`a=0`$, has degenerate centers at the origin which are limit of linear type centers of the form (7) with $`G_1=G_2=0`$ if and only if $`\mu =\lambda =0`$, or $`a=0`$. Proof: Applying the Poincaré–Liapunov method to the perturbed system (37) $$\begin{array}{ccc}\dot{x}\hfill & =\hfill & \epsilon ya(1+x)(x^44y^33y^4)+\mu y^3,\hfill \\ \dot{y}\hfill & =\hfill & \epsilon xa(1+y)(4x^3+3x^4y^4)+\lambda x^5,\hfill \end{array}$$ with $`\epsilon 0`$, we obtain the first Liapunov constant $$V_1=\frac{a\mu }{\epsilon }.$$ Vanishing $`V_1`$ we get the first center condition $`a\mu =0`$. Now, we compute the second Liapunov constant $$V_2=\frac{5a\lambda }{8\epsilon }.$$ Vanishing $`V_2`$ we obtain the second center condition $`a\lambda =0`$. So, these two conditions are necessary in order that the origin of the perturbed system (37) be a center. Therefore, these two conditions are necessary in order that the origin of system (36) be a center which is limit of linear type centers of the form (7) with $`G_1=G_2=0`$. Now we prove that these two conditions are sufficient. If $`a=0`$ we have that the system is Hamiltonian and reversible with respect to $`(x,y,t)(x,y,t)`$. Therefore, since the origin is monodromy (because it has no characteristic directions), it is a center. Now, taking the limit when $`\epsilon 0`$, we obtain a Hamiltonian system which has a degenerate center at the origin. If $`\mu =\lambda =0`$ it can be shown that system (36) has a monodromy singular point at the origin. Moreover, it has the analytic first integral $$H(x,y)=(1+x)^1(1+y)^1(x^4+y^4),$$ defined in a neighborhood of the origin. Therefore, system (36) has a degenerate center at the origin. We note that this degenerate center is neither time–reversible nor Hamiltonian. By the same arguments that for the nilpotent centers of polynomial differential systems, the degenerate centers of polynomial differential systems which are limit of linear type centers of the form (7) under the assumptions of Theorem 2, in fact, are limit of linear type centers of the form (7) where the two analytic functions $`G_1`$ and $`G_2`$ are always polynomials. ## 5. On the cyclicity of nilponent and degenerate centers Let $`p`$ be a center of a polynomial vector field of degree $`m`$. The cyclicity, $`c_n(p)`$, of $`p`$ is the maximum number of limit cycles, taking into account their multiplicity, that can bifurcate from the singular point $`p`$ when we perturb it into the class of all polynomial differential systems of degree $`nm`$. For a linear type center $`p`$ of a polynomial differential system of degree $`m`$ it is known that if the number of its independent Liapunov constants is $`k`$, then the cyclicity $`c_m(p)k1`$ if the Bautin ideal is radical, see for instance . Moreover, if we perturb a polynomial differential linear type center of degree $`m`$ and cyclicity $`k1`$ inside the class of all polynomial vector fields of degree $`m`$, we can get perturbed vector fields with exactly $`k1`$ hyperbolic limit cycles bifurcating from the center. This is due to the relationship between the Liapunov constants and the coefficients of the Poincaré map near a center. For more details on this subject see . As we have seen in the examples, in general, the Liapunov constants are not well–defined when $`\epsilon 0`$, see for instance the proof of Propositions 8 and 13. Therefore, we must impose that the limit of the Liapunov constants when $`\epsilon 0`$ be well–defined. If the Liapunov constants are well–defined when $`\epsilon 0`$, then the Poincaré map obtained by the limit $`\mathrm{\Pi }=lim_{\epsilon 0}\mathrm{\Pi }_\epsilon `$ gives a formal series which can be not convergent at any positive radius. Hence, we must also impose to the Poincaré map to be convergent in a neighborhood of the origin. Taking into account these conditions we can establish the following result. ###### Proposition 14. Suppose that the origin of a polynomial differential system (2) or (3) of degree $`m`$ is a center $`p`$, and that this system is limit when $`\epsilon 0`$ of polynomial differential systems of degree $`n`$ of the form (6) or (7), respectively, which have linear type centers $`p_\epsilon `$ at the origin. If * the Liapunov constants of a general perturbation of the same degree $`n`$ of the linear type centers (6) or (7) are well–defined when $`\epsilon 0`$, and * the limit of the Poincaré map when $`\epsilon 0`$ of the general perturbation of the same degree $`n`$ of the linear type centers (6) or (7) is analytic in a neighborhood of the origin, then, * the cyclicity $`c_n(p)`$ of the nilpotent center (2) is at most the cyclicity $`c_n(p_\epsilon )`$ of the linear type center (6) for all $`\epsilon >0`$. * the cyclicity $`c_n(p)`$ of the degenerate center (3) is at most the cyclicity $`c_n(p_\epsilon )`$ of the linear type centers (7) for $`\epsilon 0`$ sufficiently small. Proof: Let $`(P,Q)`$ be the polynomial vector field of degree $`m`$ associated to the system of the form (2) or (3) with a singular point $`p`$ of center type at the origin. By assumptions, the vector field $`(P,Q)`$ is limit when $`\epsilon 0`$ of polynomial vector fields $`(P_\epsilon ,Q_\epsilon )`$ of degree $`n`$ associated to systems of the form (6) or (7), respectively, which have linear type centers $`p_\epsilon `$ at the origin. Let $`(P_\epsilon ^{},Q_\epsilon ^{})`$ be a general perturbed polynomial vector field of degree $`n`$ of the vector field $`(P_\epsilon ,Q_\epsilon )`$. Taking the limit of $`(P_\epsilon ^{},Q_\epsilon ^{})`$ when $`\epsilon 0`$, we obtain a perturbed polynomial vector field $`(P^{},Q^{})`$ of degree at most $`n`$ of the vector field $`(P,Q)`$. Since the Liapunov constants of the system $`(P_\epsilon ^{},Q_\epsilon ^{})`$ are well–defined when $`\epsilon 0`$, and the Poincaré map $`\mathrm{\Pi }=lim_{\epsilon 0}\mathrm{\Pi }_\epsilon `$ of $`(P^{},Q^{})`$ is analytic in a neighborhood of the origin, we can control the ciclicity of the polynomial vector field $`(P^{},Q^{})`$ by the Poincaré map $`\mathrm{\Pi }`$ (with the same restrictions that for linear type centers). Moreover, the number of independent Liapunov constants of the system $`(P^{},Q^{})`$ is at most the number of the independent Liapunov constants of the system $`(P_\epsilon ^{},Q_\epsilon ^{})`$. Therefore, the cyclicity $`c_n(p)`$ of the nilpotent center (2) is at most the cyclicity $`c_n(p_\epsilon )`$ of the linear type center (6) for all $`\epsilon >0`$, and the cyclicity $`c_n(p)`$ of the degenerate center (3) is at most the cyclicity $`c_n(p_\epsilon )`$ of the linear type centers (7) for $`\epsilon 0`$ sufficiently small. Then the proposition follows. In general the degenerate problems present the more rich structure. For instance, when we look for algebraic limit cycles into the quadratic polynomial vector fields, there is one which is given by a non–degenerate algebraic curve (the algebraic limit cycle of degree 2), but there are many others (the algebraic limit cycles of degree 4, 5, 6, and perhaps others) that are given by degenerate algebraic curves. Here, a degenerate algebraic curve is an algebraic curve having singular points. For more details about algebraic limit cycles see . However, in the assumptions of Proposition 14, it is clear that the cyclicity of a non–linear type center $`p`$ which is limit of linear type centers, is not more rich than the cyclicity of the linear type centers. Hence, Proposition 4 follows from Proposition 14. ## 6. Appendix: Homogeneous and quasi–homogeneous systems In this appendix we introduce two classes of polynomial vector fields having degenerate centers. For more details about them see and . We consider polynomial differential systems in $`^2`$ of the form (38) $$\dot{x}=P(x,y),\dot{y}=Q(x,y),$$ where $`P`$ and $`Q`$ are real polynomials in the variables $`x`$ and $`y`$. We say that this system has degree $`m`$ if $`m`$ is the maximum of the degrees of $`P`$ and $`Q`$. If $`P`$ and $`Q`$ are coprime homogeneous polynomials of degree $`m`$, then the centers of systems (38) are characterized by: (i) the homogeneous polynomial $`xQ(x,y)yP(x,y)`$ has no real factors (so $`m`$ is odd), and (ii) $$_0^{2\pi }\frac{f(\theta )}{g(\theta )}𝑑\theta =0.$$ Here $`f(\theta )`$ $`=`$ $`\mathrm{cos}\theta P(\mathrm{cos}\theta ,\mathrm{sin}\theta )+\mathrm{sin}\theta Q(\mathrm{cos}\theta ,\mathrm{sin}\theta ),`$ $`g(\theta )`$ $`=`$ $`\mathrm{cos}\theta Q(\mathrm{cos}\theta ,\mathrm{sin}\theta )\mathrm{sin}\theta P(\mathrm{cos}\theta ,\mathrm{sin}\theta ).`$ Moreover, all the homogeneous centers are global centers; i.e. the periodic orbits surrounding the center fulfill all $`^2`$ . In what follows $`p`$ and $`q`$ always will denote positive integers. We say that the function $`H(x,y)`$ is $`(p,q)`$–quasi–homogeneous of weight degree $`m0`$ if $`H(ł^px,ł^qy)=ł^mH(x,y)`$ for all $`ł`$. We say that system (38) is $`(p,q)`$–quasi–homogeneous of weight degree $`m0`$ if $`P`$ and $`Q`$ are $`(p,q)`$–quasi–homogeneous functions of weight degrees $`p1+m`$ and $`q1+m`$, respectively. Note that the $`(1,1)`$–quasi–homogeneous systems of weight degree $`m`$ are the classical homogeneous polynomial differential systems of degree $`m`$. We note that if system (38) is $`(p,q)`$–quasi–homogeneous, then the differential equation $`dy/dx=Q/P`$ (another way to write system (38)) is invariant by the change of variables $`(x,y)(ł^px,ł^qy)`$. If $`P`$ and $`Q`$ are coprime, then the centers of the $`(p,q)`$–quasi–homogeneous systems (38) of degree $`m`$ are characterized by: (i) the $`(p,q)`$–quasi–homogeneous polynomial $`pxQ(x,y)qyP(x,y)`$ has no real factors, and (ii) $$_0^{2\pi }\frac{F(\theta )}{G(\theta )}𝑑\theta =0.$$ Here $`F(\theta )`$ $`=`$ $`\text{Cs }^{2q1}\theta P(\text{Cs }\theta ,\text{Sn }\theta )+\text{Sn }^{2p1}\theta Q(\text{Cs }\theta ,\text{Sn }\theta ),`$ $`G(\theta )`$ $`=`$ $`p\text{Cs }\theta Q(\text{Cs }\theta ,\text{Sn }\theta )q\text{Sn }\theta P(\text{Cs }\theta ,\text{Sn }\theta ),`$ and $`\text{Cs }\theta `$ and $`\text{Sn }\theta `$ are the $`(q,p)`$–trigonometric functions. Moreover, all the $`(p,q)`$–quasi–homogeneous centers are global centers. We recall that the $`(p,q)`$–trigonometric functions $`z(\theta )=\text{Cs }\theta `$ and $`w(\theta )=\text{Sn }\theta `$ are the solution of the following initial value problem $$\dot{z}=w^{2p1},\dot{w}=z^{2q1},z(0)=p^{\frac{1}{2q}},w(0)=0.$$ It easy to check that the functions $`\text{Cs }\theta `$ and $`\text{Sn }\theta `$ satisfy the equality $$p\text{Cs }^{2q}\theta +q\text{Sn }^{2p}\theta =1.$$ For $`p=q=1`$ we have that $`\text{Cs }\theta =\mathrm{cos}\theta `$ and $`\text{Sn }\theta =\mathrm{sin}\theta `$; i.e. the $`(1,1)`$–trigonometric functions are the classical ones. The functions $`\text{Cs }\theta `$ and $`\text{Sn }\theta `$ are $`\tau `$–periodic functions with $$\tau =2p^{\frac{1}{2q}}q^{\frac{1}{2p}}\frac{\mathrm{\Gamma }(\frac{1}{2p})\mathrm{\Gamma }(\frac{1}{2q})}{\mathrm{\Gamma }(\frac{1}{2p}+\frac{1}{2q})}.$$ Acknowledgments. The authors thank some suggestions of the referee that have improved the present paper. The second author is partially supported by a DGICYT grant number MTM2005-06098-C02-02 and by a CICYT grant number 2005SGR 00550, and by DURSI of Government of Catalonia “Distinció de la Generalitat de Catalunya per a la promoció de la recerca universitària”. The third author is partially supported by a DGICYT grant number MTM2005-06098-C02-01 and by a CICYT grant number 2005SGR 00550.
warning/0506/hep-ph0506163.html
ar5iv
text
# Avant-propos ### Avant-propos Ce cours a t donn par John Ellis en septembre 2004 l’occasion de la 36 me cole d’ t de Gif-sur-Yvette, intitul e "Le futur de la physique des hautes nergies". Le public tait principalement compos d’exp rimentateurs et le but de ce cours tait de pr senter les physiques possibles au-del du Mod le Standard les plus connues et les plus tudi es. Il a t crit par Julien Welzel (chapitres 1, 2, 3 et 4-section 1) et David Gherson (chapitre 4-sections 2, 3, 4, 5), tous deux tudiants en th se dans le groupe de physique th orique de l’Institut de Physique Nucl aire de Lyon. Nous remercions chaleureusement John Ellis et Patrick Janot pour nous avoir laiss l’opportunit de r diger le manuscrit du cours et pour avoir incit et encourag cette pratique. Nous en tirons de grands b n fices sur notre compr hension de la physique au-del du Mod le Standard. Nous remercions aussi sinc rement Aldo Deandrea et Maurice Kibler pour leurs commentaires et leurs critiques sur le cours. La bibliographie n’est pas exhaustive, nous avons pr f r donner les r f rences g n rales qui nous ont aid r diger. Cependant, il y a quand m me un certain nombre de r f rences plus sp cifiques pour citer les sources des diff rentes valeurs, graphes,… que nous avons utilis . De plus, ce cours est prendre dans le contexte de l’ cole et compl mentairement d’autres. Ainsi, il est recommand de consulter les cours de D. Treille, G. Unal, A. Blondel et A. De Roeck, par exemple, si on veut avoir de plus amples d tails sur les aspects plus exp rimentaux. J.W, D.G ## Chapitre 1 Nouvelle physique et brisure lectrofaible La n cessit d’une nouvelle physique n’est pas une observation r cente m me si jusqu’ il y a quelques ann es, le Mod le Standard (MS) de la physique des particules n’avait pas t mis en d faut exp rimentalement aupr s des acc l rateurs. La physique des neutrinos avec le ph nom ne des oscillations de neutrinos n’a fait que confirmer exp rimentalement ce que nous savions dej : le MS souffre d’insuffisances et plusieurs indices ph nom nologiques ne peuvent trouver une explication qu’au-del du MS. Nous vivons une p riode charni re pour la physique des particules car nous ne sommes qu’ quelques ann es des premiers r sultats du grand collisionneur de protons, le LHC, o de la nouvelle physique fera son apparition <sup>1</sup><sup>1</sup>1Soyons optimistes !. C’est un moment tr s important pour faire un "bilan" et un r capitulatif des possibles nouvelles physiques et des id es qui se sont d velopp es. Beaucoup de cours et livres pr sentent et exposent les d tails de la construction de ce qu’on appelle le Mod le Standard. Nous pr f rerons ici ne rappeler que le contenu physique et quelques d tails math matiques. Nous ferons aussi le choix de ne pas insister sur le succ s de cette th orie qui fait galement l’objet de nombreux expos s, pour mieux pr senter les insuffisances. Ainsi il sera plus ais de comprendre les motivations de l’ laboration d’une th orie plus compl te qui expliquerait mieux le monde que nous observons. Dans ce chapitre introductif, nous allons donc tout d’abord exposer la description actuelle de la physique des particules r sum e dans le Mod le Standard puis nous passerons en revue quelques insuffisances intrins ques. Ainsi la n cessit d’une nouvelle physique apparaitra plus clairement. Dans un second temps nous discuterons bri vement de la place des neutrinos dans le MS et de l’impact de la d couverte r cente de leurs oscillations. La troisi me partie, la plus longue, sera consacr e la brisure lectrofaible, dont l’origine est la seule inconnue du Mod le Standard. Nous verrons comment devrait intervenir le m canisme de Higgs dans la g n ration des masses et nous discuterons aussi des alternatives un boson de Higgs l mentaire que proposent les mod les de technicouleur. Puisque nous sommes l’aube des r sultats exp rimentaux du LHC nous pr senterons les possibilit s de d couverte du boson de Higgs ainsi que les mod les alternatifs dans le cas o le Higgs ne paraitrait pas dans la gamme de masses o nous l’attendons. Nous terminerons ce chapitre en pr sentant un exemple de "carte routi re" des possibles nouvelles physiques et des nigmes qu’il reste lucider dans les ann es venir. ### 1.1 Le Mod le Standard et ses probl mes #### 1.1.1 Le cadre actuel : le Mod le Standard de la physique des particules ##### Particules, interactions et sym tries Avant d’aller explorer la physique des hautes nergies et les th ories possibles, il est bon de se rem morer la vision actuelle de la physique des particules du Mod le Standard. Aujourd’hui, en physique des particules, le monde est d crit comme tant un espace 3 dimensions spatiales et 1 dimension temporelle contenant des particules l mentaires de la mati re, les fermions, qui interagissent entre elles par l’interm diaire de 4 forces. Ces 4 interactions fondamentales – lectro-magn tique, forte, faible et gravitationelle – se comprennent par l’ change d’un autre type de particules l mentaires, les bosons de jauge. On sch matise les interactions entre particules par les diagrammes de Feynman. Toutes ces interactions entre particules ob issent des sym tries locales et internes et les propri t s des particules s’en d duisent. Les th ories qui d crivent ces interactions fondamentales sont toutes des th ories de jauge. * L’ lectrodynamique quantique (QED) qui est la version relativiste et quantique de l’ lectromagn tisme est math matiquement construite autour de la sym trie de jauge locale U(1) (sym trie locale de la phase des champs). Son boson de jauge associ est le photon, de masse nulle<sup>2</sup><sup>2</sup>2Ce qui est une cons quence directe de l’invariance de jauge. Cette derni re emp che l’introduction de terme de masse $`\colorbox[rgb]{1,1,1}{$m$}\colorbox[rgb]{1,1,1}{$G$}^\colorbox[rgb]{1,1,1}{$\mu $}\colorbox[rgb]{1,1,1}{$G$}_\colorbox[rgb]{1,1,1}{$\mu $}`$ dans le lagrangien et donc d’une masse pour les champs de jauge $`\colorbox[rgb]{1,1,1}{$G$}^\colorbox[rgb]{1,1,1}{$\mu $}`$., de spin 1, et il se couple la charge lectrique des fermions. * La QCD d crit l’interaction forte et celle-ci se produit par l’ change de gluons, qui sont au nombre de huit<sup>3</sup><sup>3</sup>3 Le nombre de bosons de jauge est donn par le nombre de g n rateurs de la sym trie de jauge. Pour les th ories type SU(N), ce nombre est $`\colorbox[rgb]{1,1,1}{$N$}^\colorbox[rgb]{1,1,1}{$2$}\colorbox[rgb]{1,1,1}{$$}\colorbox[rgb]{1,1,1}{$1$}`$. et qui se couplent la charge de "couleur" des fermions. Cette charge de couleur est l’analogue "fort" de la charge lectrique bien connue. La masse des gluons est nulle<sup>4</sup><sup>4</sup>4C’est aussi une cons quence directe de l’invariance de jauge. et ils ont un spin 1. L’interaction forte se distingue entre autres par 2 propri t s importantes, le confinement et la libert asymptotique <sup>5</sup><sup>5</sup>5 La charge de couleur du quark diminue quand on se ”rapproche” du nuage de gluons autour du quark. A la limite asymptotique (distances infiniment courtes), l’interaction de couleur entre les quarks est donc nulle. Ceux-ci sont donc quasi-libres tr s courtes distances. C’est la libert asymptotique. Inversement, des distances grandes (environ $`\colorbox[rgb]{1,1,1}{$10$}^{\colorbox[rgb]{1,1,1}{$$}\colorbox[rgb]{1,1,1}{$13$}}`$ cm), l’interaction de couleur devient tr s forte et les quarks sont confin s en hadrons et n’existent pas l’ tat libre. La d monstration de la libert asymptotique a d’ailleurs valu cette ann e le prix Nobel de physique D. J. Gross, F. Wilczek et H. D .Politzer. En revanche, la d monstration du confinement reste toujours un d fi pour les th oriciens.. * Le mod le de Glashow-Weinberg-Salam (GWS) d crit la fois l’interaction faible et l’interaction lectromagn tique, r unies sous le terme d’interaction lectrofaible. Mais ce mod le n’unifie pas vraiment les 2 interactions en une seule, il les englobe math matiquement dans un m me formalisme, un m me groupe de sym trie de jauge : $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}`$. Dans ce mod le il y a en effet deux constantes de couplage diff rentes, conventionnellement $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$, respectivement. Si on veut retrouver la constante de couplage lectromagn tique $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}`$, on utilise la relation : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}{\sqrt{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (1.1) Les bosons de jauge associ s cette sym trie sont au nombre de quatre <sup>6</sup><sup>6</sup>6 Comme dans le cas de QCD, math matiquement, le nombre de bosons de jauge est li au groupe de sym trie - voir la note 3., trois "$`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$" pour $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ et un "$`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}`$" pour $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Mais les bosons physiques de la th orie, ceux qui s’observent dans les interactions, sont le photon $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}`$ et les bosons $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$. Les bosons $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$ s’obtiennent par une rotation, dans l’espace interne, des bosons $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}`$ d’un angle $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ appel angle lectrofaible : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\begin{array}{c}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\begin{array}{cc}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cos}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{sin}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{sin}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cos}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\begin{array}{c}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (1.2) et nous avons la relation : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{sin}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (1.3) qui vaut exp rimentalement $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0.23120}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0.00015}$}`$ d’apr s le PDG 2004 . C’est de ce point de vue que l’on consid re l’unification des deux interactions. La "vraie" unification (une constante de couplage unique pour l’ensemble) se produirait plut t des nergies de l’ordre de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{16}$}}`$ GeV, et incluerait aussi l’interaction forte. Nous tudierons ceci dans le cadre des th ories de Grande Unification beaucoup plus tard, au dernier chapitre. Les trois bosons $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ ont des masses approximatives de 80 et 91 GeV respectivement. La masse non-nulle des bosons faibles rend l’interaction de tr s courte port e <sup>7</sup><sup>7</sup>7Puisque $`\colorbox[rgb]{1,1,1}{$\mathrm{\Delta }$}\colorbox[rgb]{1,1,1}{$E$}\colorbox[rgb]{1,1,1}{$\mathrm{\Delta }$}\colorbox[rgb]{1,1,1}{$T$}\colorbox[rgb]{1,1,1}{$$}\colorbox[rgb]{1,1,1}{$\mathrm{}$}`$, nous trouvons que la port e $`\colorbox[rgb]{1,1,1}{$L$}\colorbox[rgb]{1,1,1}{$$}\colorbox[rgb]{1,1,1}{$\mathrm{}$}\colorbox[rgb]{1,1,1}{$/$}\colorbox[rgb]{1,1,1}{$c$}\colorbox[rgb]{1,1,1}{$M$}`$., si bien qu’ des nergies faibles devant ces masses on retrouve la th orie de Fermi des interactions faibles. Mais l’interaction faible a aussi une sp cificit tr s importante ph nom nologiquement : la violation de la parit (c’est- -dire la non-invariance par renversement des coordonn es spatiales : $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}`$) dans les interactions courants charg s (par change de bosons charg s $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$). Cette sp cificit se traduit par le fait que seules les particules de chiralit <sup>8</sup><sup>8</sup>8Toute particule de spin 1/2 peut tre d compos e en une partie appel e ”gauche” et une partie appel e ”droite” qui ont des transformations diff rentes sous le groupe de Lorentz. La chiralit se confond d’ailleurs avec l’h licit , projection du spin sur l’impulsion, dans le cas de particules de masses nulles. gauche sont sensibles l’interaction faible. * Enfin la quatri me et derni re interaction connue mais qui n’est pas comprise dans le Mod le Standard, la gravitation, est suppos e v hicul e par le graviton, de masse nulle et de spin 2. Elle ne joue pas de r le en physique des particules aux nergies qui nous sont accessibles (quelques TeV). L’interaction gravitationnelle est d crite classiquement par la Relativit G n rale que l’on peut consid rer comme une th orie de jauge bas e sur une invariance locale de l’espace-temps (le groupe de sym trie est alors le groupe de Poincar ). Les sym tries sont tr s utiles pour d crire les interactions fondamentales mais ont aussi d’importantes implications pour la classification des fermions. Ces constituants de la mati re sont class s en 2 cat gories, de propri t s diff rentes (ils ont des nombres quantiques et des transformations de sym trie diff rents). Le tableau (1.1) r sume le contenu en particules du Mod le Standard. Nous avons pour les fermions : -d’un cot les quarks, sensibles aux interactions forte, faible et lectromagn tique (et gravitationnelle). Ce sont des triplets de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}`$ qui existent donc en 3 couleurs (conventionnellement rouge, vert ou jaune, bleu). Ils se d clinent en 3 saveurs ($`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$) de charge lectrique $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}`$ et 3 saveurs ($`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$) de charge lectrique $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}`$ ainsi qu’en 2 chiralit s (gauche et droite). -d’un autre cot les leptons, qui ne sont pas sensibles l’interaction forte mais sont sensibles aux trois autres. Ils ne portent pas de charge de couleur et ont des masses g n ralement moins lev es que les quarks. Nous distinguons les leptons charg s $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}`$ ( lectron, muon, tau) qui existent avec les 2 chiralit s gauche et droite et les neutrinos de charge lectrique nulle mais uniquement gauches<sup>9</sup><sup>9</sup>9Les neutrinos droits n’ont pas t observ s l’heure actuelle, voir plus loin.. Quarks et leptons sont aussi r partis en 3 g n rations ou familles, de masses de plus en plus l v es. Pour viter les anomalies <sup>10</sup><sup>10</sup>10A l’origine, les anomalies viennent de la brisure des sym tries classiques du lagrangien par les corrections quantiques boucles. On les observe par exemple dans la d sint gration du $`\colorbox[rgb]{1,1,1}{$\pi $}^\colorbox[rgb]{1,1,1}{$0$}`$ en 2 photons. Pour les liminer, il faut introduire des conditions ext rieures comme des conditions sur les hypercharges ou sur les nombres de quarks et leptons. il faut que le nombre de g n rations soit le m me entre quarks et leptons. Ce nombre est fix 3 gr ce la mesure de la largeur du Z.<sup>11</sup><sup>11</sup>11En mesurant la largeur de d sint gration du boson $`\colorbox[rgb]{1,1,1}{$Z$}^\colorbox[rgb]{1,1,1}{$0$}`$, c’est- -dire l’inverse de son temps de vie, les exp riences du LEP ont pu d duire le nombre de saveurs de neutrinos l gers. Ils sont au nombre de $`\colorbox[rgb]{1,1,1}{$2.984$}\colorbox[rgb]{1,1,1}{$\pm $}\colorbox[rgb]{1,1,1}{$0.011$}`$. D’apr s le MS, cela limite le nombre de doublets de leptons et pour viter l’apparition d’anomalies cela limite aussi le nombre de quarks. Il n’y a donc que 3 g n rations de quarks et leptons. Les antiparticules correspondantes poss dent les m me masses, mais les charges lectriques et les chiralit s sont oppos es. Math matiquement, les particules se distinguent par leurs propri t s de transformation sous les diff rents groupes de sym tries de jauge du Mod le Standard, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}`$. A chacun de ces groupes est associ un nombre quantique: la couleur, l’isospin faible $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}`$ et sa troisi me composante $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}`$ et l’hypercharge $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}`$. La charge lectrique s’obtient par la combinaison lin aire $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$. La derni re colonne du tableau (1.1) indique, pour chacun des 3 groupes du MS, la repr sentation laquelle appartient la particule. Les nombres quantiques associ s s’en d duisent. Par exemple, l’ lectron gauche se transforme dans le MS selon la repr sentation not e (1,2,-1). Ceci signifie : -que c’est un singulet de couleur (charge de couleur nulle, il ne subit pas l’interaction forte), -qu’il forme avec le neutrino gauche un doublet d’isospin T=1/2 (et son $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}`$, la projection de l’isospin faible sur un axe, vaut -1/2 ), -qu’il poss de une hypercharge gale -1. Nous pouvons v rifier que nous obtenons bien une charge lectrique -1 en unit de la charge lectrique de l’ lectron, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}`$. ##### Le lagrangien du Mod le Standard Nous venons de d crire le contenu du MS en champs de mati re et en champs de jauge, nous pouvons maintenant r sumer tout ceci dans le lagrangien de la th orie. Mais il manque encore quelques lements essentiels qui n’ont pas t abord s. Depuis Cabibbo et gr ce Glashow, Iliopoulos, Maiani puis Kobayashi et Maskawa, on sait maintenant que les quarks sont "m lang s". C’est un des aspects les plus importants pour la ph nom nologie. On distingue les tats de saveurs qui sont les tats propres d’interaction faible et les tats de masse qui sont les tats propres de propagation. Ces deux bases ne sont pas identiques, et la matrice de Cabibbo, Kobayashi et Maskawa (CKM) est la matrice de changement de base : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\begin{array}{c}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{K}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\begin{array}{c}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\begin{array}{ccc}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\begin{array}{c}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}$$ (1.4) La matrice $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{K}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}}`$ s’ crit souvent sous la forme de trois rotations successives dans l’espace des saveurs et une phase. Cette red finition de la base des quarks n’a d’effet que sur les courants charg s ( change d’un $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$) car la matrice est unitaire ($`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$). Les courants neutres ne changent que tr s peu la saveur <sup>12</sup><sup>12</sup>12 On utilise en g n ral l’abr viation FCNC issue de l’anglais Flavour Changing Neutral Current pour parler de courant neutre qui change la saveur.. La matrice CKM est complexe ce qui se traduit par des phases $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}}`$ dans la d finition des tats propres. Mais nous pouvons montrer qu’une seule phase est physique (observable) et reste a priori non-nulle. Physiquement, c’est cette phase qui serait l’origine de la violation de la sym trie $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}`$ dans certaines interactions faibles <sup>13</sup><sup>13</sup>13Par exemple dans $`\colorbox[rgb]{1,1,1}{$K$}\colorbox[rgb]{1,1,1}{$$}\colorbox[rgb]{1,1,1}{$\pi $}\colorbox[rgb]{1,1,1}{$\pi $}`$ ou dans $`\colorbox[rgb]{1,1,1}{$B$}^\colorbox[rgb]{1,1,1}{$0$}\colorbox[rgb]{1,1,1}{$$}\colorbox[rgb]{1,1,1}{$K$}_\colorbox[rgb]{1,1,1}{$S$}\colorbox[rgb]{1,1,1}{$J$}\colorbox[rgb]{1,1,1}{$/$}\colorbox[rgb]{1,1,1}{$\psi $}`$. Si la sym trie $`\colorbox[rgb]{1,1,1}{$C$}\colorbox[rgb]{1,1,1}{$P$}`$ n’est pas respect e, alors un processus r alis avec les antiparticules ne donnera pas le m me r sultat (c’est- -dire la m me section efficace, par exemple) que le m me processus avec des particules.. Le lagrangien du MS se construit en suivant deux r gles essentielles : l’invariance de jauge sous le groupe de sym trie du MS et la renormalisabilit . En effet, cette derni re assure que les quantit s mesurables calcul es seront finies car les infinit s rencontr es ne sont pas physiques. Ceci rend ce mod le tr s pr dictif contrairement l’ancienne th orie des interactions faibles, la th orie de Fermi. Le lagrangien (1.5) comporte 4 parties, une pour chaque ligne; la premi re d crit le secteur de jauge c’est- -dire toute la dynamique des champs de jauge (des bosons de jauge). Le deuxi me terme est le secteur de Dirac dans lequel nous retrouvons tous les champs de mati re $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}`$ (les fermions) et leurs interactions avec les champs de jauge. Le troisi me terme est le secteur de Yukawa qui contient les interactions des fermions avec le champ de Higgs $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}`$ et qui donne apr s brisure de la sym trie lectrofaible toutes les masses des fermions. Enfin, la derni re ligne correspond au secteur de Higgs qui r alise le m canisme de Higgs dont nous parlerons plus loin. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}`$ (1.5) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{D}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{D}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}`$ #### 1.1.2 Le Mod le Standard et les tests de pr cision lectrofaibles M me si ce cours est orient vers les nouvelles physiques et l’apr s-MS, ne pas pr senter l’extraordinaire accord du MS avec les donn es exp rimentales ne serait pas lui rendre justice. C’est en effet un mod le tr s pr dictif et qui a t test et v rifi avec une pr cision sup rieure au pourcent, dans une large vari t d’exp riences, et dans un vaste domaine d’ nergie : de la violation de la parit dans les atomes avec une (impulsion)<sup>2</sup> transf r e, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$, d’environ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}}`$ GeV<sup>2</sup> des collisions proton-antiproton $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}`$ GeV<sup>2</sup>. Il existe des r f rences r centes dans lesquelles tous ces tests sont pr sent s, nous nous contenterons de r sumer l’accord de ce mod le avec les nombreux tests de pr cision lectrofaibles dans le tableau (1.2). On y retrouve par exemple les r sultats de la mesure de la largeur du $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$, de la masse du $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ et de celle du top. L’ensemble des tests forme un ensemble de contraintes tr s fortes. A l’examen du tableau (1.2) nous ne pouvons pas encore dire que nous avons vu de d viation significative des pr dictions du MS et donc de nouvelle physique. N anmoins, cet accord contraint fortement les nouvelles physiques possibles. Jusqu’ pr sent le Mod le Standard est donc une description tr s performante de la r alit des chelles d’ nergies inf rieures quelques centaines de GeV. #### 1.1.3 Les insuffisances du Mod le Standard Le constat exp rimental pr c dent, en faveur du MS, confront au constat th orique suivant est source d’ tonnements dans la communaut des physiciens des particules. En effet, le MS n’explique pas, par exemple, les nombres quantiques des particules (charge lectrique, isospin faible, couleur,…), ne pr dit pas le spectre de masse ni n’inclut l’interaction gravitationnelle. Passons un peu de temps sur chaque point pour mieux comprendre pourquoi le Mod le Standard ne peut pas tre la description ultime de la nature. ##### Les param tres libres Outre les nombres quantiques des particules, le MS contient au moins 19 param tres libres, non d termin s par la th orie et ajust s a posteriori par l’exp rience. Il n’explique donc pas l’origine de la valeur de ces param tres. Si nous faisons le d compte, nous avons : \- 3 constantes de couplage pour chacune des 3 interactions (le MS ne nous dit pas pourquoi l’interaction faible est si faible ni pourquoi l’interaction forte est si forte basse nergie, les valeurs ont t d duites des mesures). De fa on quivalente, on peut remplacer ces trois param tres libres par $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{sin}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$. \- 1 param tre de QCD, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{D}$}}`$ qui correspond une possible violation de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}`$ dans les interactions fortes <sup>14</sup><sup>14</sup>14Ce param tre est la cause de ce que l’on appelle le ”strong $`\colorbox[rgb]{1,1,1}{$C$}\colorbox[rgb]{1,1,1}{$P$}`$ problem” car on ne comprend pas pourquoi sa valeur est si faible voire peut- tre m me nulle. En effet, les mesures exp rimentales donnent une limite sup rieure sur $`\colorbox[rgb]{1,1,1}{$\theta $}_{\colorbox[rgb]{1,1,1}{$Q$}\colorbox[rgb]{1,1,1}{$C$}\colorbox[rgb]{1,1,1}{$D$}}`$ de $`\colorbox[rgb]{1,1,1}{$10$}^\colorbox[rgb]{1,1,1}{$$}\colorbox[rgb]{1,1,1}{$9$}`$ . \- les 6 masses des 6 saveurs de quarks $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ \- 3 angles de m langes des quarks + 1 phase de violation de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}`$ faible \- les 3 masses des 3 leptons charg s e, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}`$ \- les 2 param tres du potentiel de Higgs, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}`$ (ou de fa on quivalente $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ ou $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$). La d couverte r cente des oscillations de neutrinos ajoute encore 9 param tres suppl mentaires. De plus, nous avons aussi besoin de param tres libres suppl mentaires pour g n rer les masses des neutrinos par le m canisme de Seesaw <sup>15</sup><sup>15</sup>15Tout ce qui est li la masse des neutrinos sera pr cis au chapitre 3.. Ceci est un probl me que la plupart des mod les rencontrent mais parmi les param tres pr c dents certains trouvent une explication dans les Th ories de Grande Unification (les couplages de jauge, les angles de m langes) ou la supersym trie (les param tres de la brisure lectrofaible). ##### La gravitation Le Mod le Standard ne donne pas de description quantique de la th orie de la gravit . En fait, l’interaction gravitationnelle est ajout e, juxtapos e au MS sous sa forme actuelle c’est- -dire la Relativit G n rale (RG). Cela ne pose pas de probl me l’ chelle d’ nergie qui nous est accessible actuellement puisque la force gravitationelle y est consid rablement plus faible que l’interaction faible. Mais pour des nergies autour de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{19}$}}`$ GeV, elle est de m me intensit et nous ne pouvons plus la traiter classiquement via la RG. De plus, certains ph nom nes cosmologiques sont insuffisamment expliqu s et il faut introduire d’autres param tres libres, non pr dits par la th orie, comme la constante cosmologique $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}`$, au moins un param tre d’inflation, l’asym trie baryonique <sup>16</sup><sup>16</sup>16C’est- -dire l’asym trie entre le nombre de particules et d’antiparticules.,… La quantification de la gravitation, i.e. sa version quantique, est depuis longtemps une des pr occupations principales des physiciens th oriciens. La difficult vient notamment de sa non-renormalisabilit . Comme nous le verrons la fin de ce cours sur les nouvelles physiques, la th orie des supercordes semble pouvoir tre la solution et l’espoir est permis. ##### La mati re noire Au d but des ann es 90, les astronomes ont fait une d couverte <sup>17</sup><sup>17</sup>17C’est en comparant la vitesse de rotation de la galaxie avec la valeur qu’elle aurait si elle n’ tait faite que de mati re ordinaire. tr s importante pour les physiciens des particules : l’existence de mati re noire. Ils ont observ que la densit de mati re connue (baryonique) ne repr sentait que moins de 4$`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\%}$}`$ de la densit totale de l’Univers. 20$`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\%}$}`$ de la densit totale viendrait d’une mati re non baryonique qui n’existe pas dans le MS, ce que l’on appelle la mati re noire. Pire encore, le reste nous est totalement inconnu et on l’attribue une sorte d’nergie noire <sup>18</sup><sup>18</sup>18Elle est appel e ainsi car elle elle n’est pas due la mati re.. Tous les mod les au-del du MS contiennent des particules nouvelles et celles-ci, condition d’ tre stables, neutres et sans interactions fortes, peuvent jouer le r le de mati re noire. La supersym trie offre par exemple un candidat tr s s rieux, le neutralino. Par contre, l’ nergie noire est aujourd’hui un myst re total et son explication ne peut se trouver que dans les mod les qui incluent la gravitation. ##### La hi rarchie des chelles Si nous regardons le domaine des nergies, trac sur la figure (1.1), il s’ tend jusqu’ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{19}$}}`$ GeV o la gravit n’est plus n gligeable dans les interactions entre particules. Une question se pose alors, pourquoi y a-t-il 17 ordres de grandeurs qui s parent l’ chelle lectrofaible de l’ chelle de Planck ?! L non plus le MS ne donne pas de r ponse, et c’est m me l’un des plus grands d fis de la physique moderne. Peu de mod les apportent une explication satisfaisante. A plus petite chelle, la m me observation peut tre faite propos des masses de fermions connus. Pourquoi le top est il $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\times }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}`$ fois plus lourd que l’ lectron ? Il faut l aussi comprendre les raisons de cette autre hi rarchie (appel e parfois hi rarchie des saveurs). L’introduction de dimensions spatiales suppl mentaires pourrait donner une r ponse au probl me mais ce n’est souvent qu’une reformulation qu’elles proposent. Clairement, il reste des myst res qui n cessitent un autre mod le qui prendra le relais du Mod le Standard pour la physique plus haute nergie. ### 1.2 Introduction au secteur des neutrinos Cette section a pour but d’exposer ce que la plupart des physiciens consid rent comme le premier indice de l’existence d’une nouvelle physique au-del du MS. Mais avant tout, rappellons-nous les particularit s des neutrinos et leur place dans le MS. #### 1.2.1 Les neutrinos dans le Mod le Standard Le neutrino f t la premi re fois introduit par W.Pauli pour rendre compte du spectre continu de l’ lectron dans les d sint grations $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$ telles que : $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}`$ et c’est 25 ann es plus tard que F.Reines et C.Cowan Jr observ rent pour la premi re fois ces neutrinos <sup>19</sup><sup>19</sup>19En r alit c’ tait des antineutrinos lectroniques., produits par un r acteur nucl aire. Depuis les ann es 60, des neutrinos provenant du Soleil, de l’atmosph re et m me d’une supernova du Nuage de Magellan ont t d tect s. Dans la théorie électrofaible il est rangé avec l’ lectron <sup>20</sup><sup>20</sup>20Plus rigoureusement, c’est le neutrino lectronique qui est rang avec l’ lectron. Comme tous les fermions du MS, le neutrino existe en 3 familles : le neutrino lectronique, muonique, tauique. dans un doublet leptonique de chiralit gauche, d’isospin faible 1/2 et d’hypercharge $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$; voir tableau (1.1) : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\begin{array}{c}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (1.6) Il ne subit que l’interaction faible. Mais contrairement l’ lectron ou n’importe quel autre lepton charg , jusqu’ pr sent seul des neutrinos de chiralit gauche ont t observ s <sup>21</sup><sup>21</sup>21 Les neutrinos droits n’existent donc pas ? Si un neutrino droit existe, tant de chiralit droite et l’interaction faible ne concernant que les particules gauches, il n’est pas sujet cette interaction. Il ne lui reste que des interactions indirectes (nous ne comptons pas la gravitation) pour se manifester (voir chapitre 3). (ainsi que des antineutrinos droits). La faiblesse de ses interactions avec la mati re rend difficile sa d tection et les exp riences qui visent connaitre les propri t s des neutrinos atteignent donc des tailles colossales. Pour exemple, l’exp rience SuperKamiokande au Japon est constitu e d’un immense r servoir d’eau de 45 000 tonnes. Elle est capable notamment de détecter en temps réel les électrons de l’eau diffusés par interaction élastique avec les neutrinos solaires : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ L’explication de la faiblesse des interactions se trouve dans le fait que les neutrinos n’interagissent que par change de boson $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ ou $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$, c’est- -dire : \- par courant chargé (échange d’un boson vecteur $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$): $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}}{\sqrt{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (1.7) \- par courant neutre (échange d’un boson vecteur Z<sup>o</sup> ): $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cos}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (1.8) Les sections efficaces contiennent alors un facteur tr s petit: $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (1.9) Nous allons maintenant discuter de la masse des neutrinos. Les fermions, comme l’ lectron, ont un terme de masse dans le lagrangien qui s’ crit : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{D}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (1.10) C’est un terme qui lie un fermion gauche avec un fermion droit. La non-existence de neutrinos droits interdit donc ce terme de masse et nous disons que le neutrino n’a pas de masse de Dirac. Mais pouvons-nous crire un terme de masse pour $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}`$ m me sans $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$? Oui, un terme de masse de Majorana : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}$$ (1.11) avec $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}}`$. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}`$ est la matrice de conjugaison de charge ($`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$). Dans ce cas, le neutrino est sa propre antiparticule ! Or, il existe dans le MS une loi de conservation qui interdit ce terme. En effet, pour expliquer la non-observation de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}`$ <sup>22</sup><sup>22</sup>22Les limites sup rieures sur les taux de branchements c’est- -dire sur les largeurs partielles de d sint gration sont : $`\colorbox[rgb]{1,1,1}{$B$}\colorbox[rgb]{1,1,1}{$R$}\colorbox[rgb]{1,1,1}{$($}\colorbox[rgb]{1,1,1}{$\mu $}\colorbox[rgb]{1,1,1}{$$}\colorbox[rgb]{1,1,1}{$e$}\colorbox[rgb]{1,1,1}{$\gamma $}\colorbox[rgb]{1,1,1}{$)$}\colorbox[rgb]{1,1,1}{$<$}\colorbox[rgb]{1,1,1}{$1.2$}\colorbox[rgb]{1,1,1}{$\times $}\colorbox[rgb]{1,1,1}{$10$}^{\colorbox[rgb]{1,1,1}{$$}\colorbox[rgb]{1,1,1}{$11$}}`$, $`\colorbox[rgb]{1,1,1}{$B$}\colorbox[rgb]{1,1,1}{$R$}\colorbox[rgb]{1,1,1}{$($}\colorbox[rgb]{1,1,1}{$\tau $}\colorbox[rgb]{1,1,1}{$$}\colorbox[rgb]{1,1,1}{$e$}\colorbox[rgb]{1,1,1}{$\gamma $}\colorbox[rgb]{1,1,1}{$)$}\colorbox[rgb]{1,1,1}{$<$}\colorbox[rgb]{1,1,1}{$2.7$}\colorbox[rgb]{1,1,1}{$\times $}\colorbox[rgb]{1,1,1}{$10$}^\colorbox[rgb]{1,1,1}{$$}\colorbox[rgb]{1,1,1}{$6$}`$, $`\colorbox[rgb]{1,1,1}{$B$}\colorbox[rgb]{1,1,1}{$R$}\colorbox[rgb]{1,1,1}{$($}\colorbox[rgb]{1,1,1}{$\tau $}\colorbox[rgb]{1,1,1}{$$}\colorbox[rgb]{1,1,1}{$\mu $}\colorbox[rgb]{1,1,1}{$\gamma $}\colorbox[rgb]{1,1,1}{$)$}\colorbox[rgb]{1,1,1}{$<$}\colorbox[rgb]{1,1,1}{$1.1$}\colorbox[rgb]{1,1,1}{$\times $}\colorbox[rgb]{1,1,1}{$10$}^\colorbox[rgb]{1,1,1}{$$}\colorbox[rgb]{1,1,1}{$6$}`$. on introduit un nombre leptonique $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}`$ pour chaque famille et qui se conserve. Ainsi on a les nombres leptoniques suivant pour le muon $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$, l’ lectron $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ et par exemple la d sint gration $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}`$ ne serait pas permise cause de la conservation de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}`$. Ce nombre leptonique est reli une sym trie globale du lagrangien dont l’origine profonde reste tre clarifi e. Un terme de masse de Majorana comme l’ quation (1.11) a $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$, le nombre leptonique individuel n’est pas conserv . Il n’y a pas d’autre terme de masse possible, par cons quent le neutrino a une masse nulle dans le Mod le Standard ! Ceci est assez surprenant pour une particule de mati re. Mais est-ce coh rent avec les r sultats exp rimentaux ? Les observations directes sur la masse des neutrinos et les limites cosmologiques ne donnent malheureusement que des limites sup rieures, mais il y a un fait exp rimental qui entre en d saccord direct avec des masses nulles pour les neutrinos : l’observation des oscillations entre saveurs de neutrinos. #### 1.2.2 Les oscillations de neutrinos : nouvelle physique Les oscillations de neutrinos reposent sur le m me principe que le m lange CKM des quarks "down" : les tats propres de masses, i.e. de propagation, ne co ncident pas avec les tats propres de saveurs. En d’autres mots, les neutrinos produits dans les interactions faibles ( tats de saveurs) sont des combinaisons lin aires des tats de propagation. Un neutrino de saveur purement lectronique, initialement produit dans le Soleil, a donc une probabilit non-nulle de devenir purement de saveur muonique l’endroit o on le d tecte <sup>23</sup><sup>23</sup>23Dans le cas deux saveurs, nous pouvons faire l’analogie avec un syst me deux niveaux en m canique quantique dans lequel la probabilit de pr sence oscille entre les 2 niveaux la pulsation de Rabi.. Ainsi, les exp riences comme SuperKamiokande qui peuvent faire la comptabilit des neutrinos lectroniques provenants du Soleil observent un d ficit important (plus de 50$`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\%}$}`$ !) par rapport au nombre de neutrinos solaires pr dits par le mod le solaire standard <sup>24</sup><sup>24</sup>24Ce mod le est tr s performant et ne peut tre raisonnablement mis en doute.. Toutes les interpr tations autres que les oscillations, comme une d sint gration du neutrino, sont tr s d favoris s . Le ph nom ne d’oscillation entre saveurs n cessite une masse non-nulle pour les neutrinos. De plus le nombre leptonique individuel n’est pas conserv au cours du temps. Voici pourquoi nous disons que les oscillations, et donc les masses des neutrinos, demandent une physique au-del du Mod le Standard. Nous reviendrons sur ce point au chapitre 3 o les choses apparaitront plus clairement. Il est vrai que nous pouvons toujours modifier et tendre le MS pour inclure ce ph nom ne mais l’explication n’est pas naturelle. Le neutrino est visiblement un fermion part et l’origine de sa particuliarit doit se comprendre partir d’une physique plus haute nergie (par exemple dans les th ories de Grande Unification que nous verrons en d tail au dernier chapitre). ### 1.3 Le secteur de Higgs : la brisure lectrofaible #### 1.3.1 La g n ration des masses Le secteur lectrofaible du MS est d crit par la sym trie de jauge $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}`$. Or l’invariance du lagrangien sous cette sym trie ne permet pas l’introduction de termes de masse pour les bosons ou les fermions ! A ce stade de la description, aucune particule n’est donc massive ce qui n’est pas conforme nos observations. La sym trie doit donc tre bris e d’une fa on ou d’une autre. Cependant, il faut garder cette sym trie dans les interactions c’est- -dire dans le lagrangien. La brisure ne peut alors pas tre explicite, elle est spontan e. Il existe un th or me, le th or me de Goldstone qui donne les deux cons quences importantes de la brisure spontan e d’une sym trie globale : * Le lagrangien reste invariant mais l’ tat d’ nergie minimale, i.e. le vide, n’est pas invariant. La sym trie n’est alors plus manifeste dans le spectre des tats <sup>25</sup><sup>25</sup>25Car les tats excit s s’obtiennent par l’action des diff rents g n rateurs sur le vide (l’ tat fondamental) et ne poss dent plus la sym trie.. * Il existe un certain nombre d’ tats physiques de masse nulle dont les propri t s sont reli es celles des g n rateurs de la sym trie bris e. Ce sont les modes ou bosons de Goldstone. Il y en a autant que le nombre de g n rateurs bris s de la sym trie. Quand la sym trie est approch e, par exemple la sym trie chirale des quarks de l’interaction forte, les bosons de Goldstone ont une masse non-nulle, mais faible, dont l’ordre de grandeur est reli e la validit de l’approximation. Les pions, qui ont une masse d’environ 135 MeV (donc beaucoup plus faible que les masses des autres hadrons, objets typiques de l’interaction forte), jouent ce r le. On les appelle les bosons de pseudo-Goldstone de la sym trie chirale de l’interaction forte. Mais dans le Mod le Standard, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}`$ est une sym trie de jauge donc locale. Les cons quences de la brisure spontan e sont alors diff rentes : certains bosons de Goldstone disparaissent au profit de nouveaux tats de polarisation longitudinale des bosons de jauge, devenus massifs <sup>26</sup><sup>26</sup>26Nous passons de deux tats de polarisation d’un champ de jauge de spin 1 sans masse trois.. Les bosons de Goldstone restants sont aussi devenus massifs. C’est le fameux m canisme de Higgs de Brout, Englert, Higgs et Kibble et c’est le seul m canisme connu de g n ration des masses qui conserve la renormalisabilit du MS. Nous allons maintenant expliciter un peu plus ce m canisme. Prenons le cas simple d’un champ scalaire complexe $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}`$ dans une th orie de jauge $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Le lagrangien est : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{D}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{D}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{[}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{]}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (1.12) Le terme $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{[}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{]}$}`$ est le potentiel effectif et renormalisable cr par le champ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}`$, il peut s’ crire : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{[}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{]}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}$$ (1.13) avec $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$. Le champ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}`$ est appel le champ de Higgs et si nous le param trisons en fonction de deux champs r els, alors $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\sqrt{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ et le potentiel s’ crira : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{[}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{]}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}$$ (1.14) Il est manifestement invariant sous une "rotation" des champs $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ et poss de la sym trie $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{O}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ <sup>27</sup><sup>27</sup>27Le potentiel reste invariant quand on red finit les champs ainsi: $$\colorbox[rgb]{1,1,1}{$\left($}\begin{array}{c}\colorbox[rgb]{1,1,1}{$\varphi $}_\colorbox[rgb]{1,1,1}{$1$}^{^{\colorbox[rgb]{1,1,1}{$$}}}\\ \colorbox[rgb]{1,1,1}{$\varphi $}_\colorbox[rgb]{1,1,1}{$2$}^{^{\colorbox[rgb]{1,1,1}{$$}}}\end{array}\colorbox[rgb]{1,1,1}{$\right)$}\colorbox[rgb]{1,1,1}{$=$}\colorbox[rgb]{1,1,1}{$\left($}\begin{array}{cc}\colorbox[rgb]{1,1,1}{$\mathrm{cos}$}\colorbox[rgb]{1,1,1}{$\theta $}\colorbox[rgb]{1,1,1}{$e$}^{\colorbox[rgb]{1,1,1}{$i$}\colorbox[rgb]{1,1,1}{$\delta $}}& \colorbox[rgb]{1,1,1}{$\mathrm{sin}$}\colorbox[rgb]{1,1,1}{$\theta $}\\ \colorbox[rgb]{1,1,1}{$$}\colorbox[rgb]{1,1,1}{$\mathrm{sin}$}\colorbox[rgb]{1,1,1}{$\theta $}& \colorbox[rgb]{1,1,1}{$\mathrm{cos}$}\colorbox[rgb]{1,1,1}{$\theta $}\colorbox[rgb]{1,1,1}{$e$}^{\colorbox[rgb]{1,1,1}{$i$}\colorbox[rgb]{1,1,1}{$\delta $}}\end{array}\colorbox[rgb]{1,1,1}{$\right)$}\colorbox[rgb]{1,1,1}{$\left($}\begin{array}{c}\colorbox[rgb]{1,1,1}{$\varphi $}_\colorbox[rgb]{1,1,1}{$1$}\\ \colorbox[rgb]{1,1,1}{$\varphi $}_\colorbox[rgb]{1,1,1}{$2$}\end{array}\colorbox[rgb]{1,1,1}{$\right)$}$$ . Le potentiel admet une infinit de minima qui se distinguent les uns des autres par rotation : l’ tat de vide est d g n r . Quand nous choisissons un minimum particulier pour le vide, en $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\sqrt{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}`$ : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\sqrt{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\begin{array}{c}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}\end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}$$ (1.15) avec $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\sqrt{\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (1.16) nous brisons la symétrie. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ n’est plus invariant sous une transformation de jauge mais les lois dynamiques sont pr serv es : $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ reste toujours invariant de jauge. C’est tout le principe d’une brisure spontan e de sym trie, par opposition la brisure explicite o $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ perd son invariance. Si nous red finissons le champ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ comme une fluctuation autour de ce minimum particulier, en passant en coordonnées polaires nous avons : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\sqrt{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}$$ (1.17) o $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ est le degr de libert associ la sym trie de rotation du potentiel. Il correspond des modes d’excitation de masse nulle, c’est le boson de Goldstone tandis que $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ est le boson associ aux excitations radiales. En d veloppant la partie cin tique du lagrangien, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{D}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$, et en passant dans la jauge unitaire <sup>28</sup><sup>28</sup>28Notre lagrangien est par construction invariant de jauge donc nous pouvons faire ce choix particulier. Il a la particularit de faire disparaitre les bosons de Goldstone non physiques (de masse nulle) au profit de masse pour les bosons de jauge. Mais ce choix de jauge unitaire n’est pas toujours le plus pratique cause de l’existence de bosons massifs. En effet, le propagateur d’un boson massif s’ crit : $$\colorbox[rgb]{1,1,1}{$\mathrm{\Delta }$}^{\colorbox[rgb]{1,1,1}{$\mu $}\colorbox[rgb]{1,1,1}{$\nu $}}\colorbox[rgb]{1,1,1}{$=$}\colorbox[rgb]{1,1,1}{$i$}\colorbox[rgb]{1,1,1}{$\left($}\frac{\colorbox[rgb]{1,1,1}{$g$}^{\colorbox[rgb]{1,1,1}{$\mu $}\colorbox[rgb]{1,1,1}{$\nu $}}\colorbox[rgb]{1,1,1}{$$}\frac{\colorbox[rgb]{1,1,1}{$q$}^\colorbox[rgb]{1,1,1}{$\mu $}\colorbox[rgb]{1,1,1}{$q$}^\colorbox[rgb]{1,1,1}{$\nu $}}{\colorbox[rgb]{1,1,1}{$M$}^\colorbox[rgb]{1,1,1}{$2$}}}{\colorbox[rgb]{1,1,1}{$q$}^\colorbox[rgb]{1,1,1}{$2$}\colorbox[rgb]{1,1,1}{$$}\colorbox[rgb]{1,1,1}{$M$}^\colorbox[rgb]{1,1,1}{$2$}}\colorbox[rgb]{1,1,1}{$\right)$}$$ Il ne converge pas vers 0 quand $`\colorbox[rgb]{1,1,1}{$q$}^\colorbox[rgb]{1,1,1}{$\mu $}\colorbox[rgb]{1,1,1}{$$}\colorbox[rgb]{1,1,1}{$+$}\colorbox[rgb]{1,1,1}{$\mathrm{}$}`$ ( cause du terme $`\frac{\colorbox[rgb]{1,1,1}{$q$}^\colorbox[rgb]{1,1,1}{$\mu $}\colorbox[rgb]{1,1,1}{$q$}^\colorbox[rgb]{1,1,1}{$\nu $}}{\colorbox[rgb]{1,1,1}{$M$}^\colorbox[rgb]{1,1,1}{$2$}}`$ ) ce qui pose probl me pour la renormalisation. Pour y remédier, le th or me d’ quivalence nous dit que la voie suivre et de ne pas calculer les amplitudes dans la jauge unitaire et donc de garder explicitement les bosons de Goldstone., nous observons l’apparition de termes de masse, de termes d’interactions, ainsi que la disparition du champ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Le champ de jauge a maintenant une masse $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}`$ <sup>29</sup><sup>29</sup>29$`\colorbox[rgb]{1,1,1}{$e$}`$ est la constante de couplage de la th orie que nous consid rons dans cet exemple. et le champ scalaire r el $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ une masse $`\sqrt{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}`$. Comment r aliser ce m canisme dans le MS ? Nous souhaitons briser $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}`$ en $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}}`$ sans toucher $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}`$. Il nous faut donc un champ scalaire neutre de couleur, d’isospin et hypercharge faibles non nuls et au moins une composante de ce champ doit tre nulle lectriquement <sup>30</sup><sup>30</sup>30C’est pour qu’il ne se couple pas au photon et donc qu’il laisse celui-ci sans masse.. De plus, dans un mod le minimal, le m me champ de Higgs brise la sym trie lectrofaible et donne aussi une masse aux quarks et aux leptons. Dans le MS, les fermions gauches sont des doublets de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}`$ et les fermions droits des singulets, donc le produit est un doublet de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}`$. Le terme de masse des fermions, qui doit tre un scalaire sous $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}`$ <sup>31</sup><sup>31</sup>31Il doit tre invariant car le lagrangien l’est., ne peut s’obtenir que par couplage direct un champ de Higgs doublet de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}`$ <sup>32</sup><sup>32</sup>32D’apr s la th orie des groupes, $`\colorbox[rgb]{1,1,1}{$($}\colorbox[rgb]{1,1,1}{$2$}\colorbox[rgb]{1,1,1}{$$}\colorbox[rgb]{1,1,1}{$1$}\colorbox[rgb]{1,1,1}{$)$}\colorbox[rgb]{1,1,1}{$$}\colorbox[rgb]{1,1,1}{$2$}\colorbox[rgb]{1,1,1}{$=$}\colorbox[rgb]{1,1,1}{$3$}\colorbox[rgb]{1,1,1}{$$}\colorbox[rgb]{1,1,1}{$1$}`$ donc on peut former un scalaire de $`\colorbox[rgb]{1,1,1}{$S$}\colorbox[rgb]{1,1,1}{$U$}\colorbox[rgb]{1,1,1}{$($}\colorbox[rgb]{1,1,1}{$2$}\colorbox[rgb]{1,1,1}{$)$}`$ avec 2 doublets et un singulet.. Par exemple dans le cas des leptons, $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}$$ (1.18) o $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}}`$ est une constante (matrice $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\times }$}`$3 dans l’espace des saveurs) qui caract rise le couplage au champ de Higgs. Elle est appel e couplage de Yukawa et est diff rente pour chaque fermion (car chaque fermion a bien une masse distincte des autres). Pour tre en plus invariant sous $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}`$, il faut que les hypercharges v rifient la condition <sup>33</sup><sup>33</sup>33La trace de $`\colorbox[rgb]{1,1,1}{$Y$}`$ doit tre nulle car c’est un des g n rateurs diagonaux du groupe de sym trie. : $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$, donc l’hypercharge de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}`$ est 1. Le champ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}`$ s’ crit alors <sup>34</sup><sup>34</sup>34En utilisant la relation $`\colorbox[rgb]{1,1,1}{$Q$}\colorbox[rgb]{1,1,1}{$=$}\colorbox[rgb]{1,1,1}{$T$}_\colorbox[rgb]{1,1,1}{$3$}\colorbox[rgb]{1,1,1}{$+$}\frac{\colorbox[rgb]{1,1,1}{$Y$}}{\colorbox[rgb]{1,1,1}{$2$}}`$. : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\begin{array}{c}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}$$ (1.19) et il appartient la repr sentation (1,2,+1) du MS. Le potentiel effectif cr par $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}`$ est semblable celui de l’exemple pr c dent. Sa forme est la plus g n rale possible et respecte les conditions d’invariance sous $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}`$ et de renormalisabilit . Apr s brisure spontan e, dans la jauge unitaire, nous rempla ons $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}`$ par sa valeur dans le vide : $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\sqrt{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\begin{array}{c}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}\end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}`$. * Les fermions ont alors une masse $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}}{\sqrt{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}}}`$. * Les bosons de jauge ont obtenu une masse : $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cos}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}}}`$. De $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}}`$, nous sommes pass s de 4 g n rateurs de sym trie 1. D’apr s ce qui a t dit pr c demment, 3 g n rateurs sont bris s et deviennent les 3 degr s de polarisation longitudinale des $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ qui sont maintenant massiques. Or sur les 4 champs scalaires r els <sup>35</sup><sup>35</sup>35Un doublet de champs scalaires complexes peut s’ crire en fonction de quatre champs scalaires r els : $$\colorbox[rgb]{1,1,1}{$\varphi $}\colorbox[rgb]{1,1,1}{$=$}\colorbox[rgb]{1,1,1}{$\left($}\begin{array}{c}\colorbox[rgb]{1,1,1}{$\varphi $}_\colorbox[rgb]{1,1,1}{$1$}\colorbox[rgb]{1,1,1}{$+$}\colorbox[rgb]{1,1,1}{$i$}\colorbox[rgb]{1,1,1}{$\varphi $}_\colorbox[rgb]{1,1,1}{$2$}\\ \colorbox[rgb]{1,1,1}{$\varphi $}_\colorbox[rgb]{1,1,1}{$3$}\colorbox[rgb]{1,1,1}{$+$}\colorbox[rgb]{1,1,1}{$i$}\colorbox[rgb]{1,1,1}{$\varphi $}_\colorbox[rgb]{1,1,1}{$4$}\end{array}\colorbox[rgb]{1,1,1}{$\right)$}$$ qui composent $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}`$, 3 sont les bosons de Golstones absorb s, il reste donc un boson de Goldstone physique, le boson de Higgs, avec une masse : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (1.20) Si nous r sumons ce qui vient d’ tre dit, le m canisme de Higgs permet de briser spontan ment la sym trie lectrofaible, donnant alors une masse aux bosons de jauge $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$. Les fermions obtiennent eux une masse en se couplant avec le champ de Higgs et ce couplage est proportionnel $`{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}}}`$. De cette brisure, il reste le boson de Higgs, de spin nul et de masse $`\sqrt{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}`$. A partir de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ ou de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}`$, reli s par la relation $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\sqrt{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{G}}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{sin}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (1.21) on peut d duire la valeur de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}`$ : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\sqrt{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{246}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{GeV}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (1.22) Connaissant $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$, la masse du boson de Higgs est donc le seul param tre libre de ce secteur ce qui rend possible la pr diction de la production et des d sint grations du Higgs en fonction de sa masse. Dans la figure (1.2) nous montrons les d sint grations du Higgs les plus importantes en fonction de sa masse. Notamment, pour $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$ entre 100 et 150 GeV, les d sint grations en une paire $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}`$ ou $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ semblent tr s prometteuses et les collaborations CDF et D$`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ en font la recherche tr s active. Le LHC pourra davantage explorer la ph nom nologie du Higgs. #### 1.3.2 Les propri t s du Higgs et les mod les Technicouleur Si nous calculons les corrections radiatives <sup>36</sup><sup>36</sup>36Ces corrections sont dues aux boucles de particules virtuelles. la masse du Higgs dans le MS, nous observons qu’elles divergent quadratiquement : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒪}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pi }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}$$ (1.23) Quand le cut-off $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}`$ est sup rieur au TeV, les corrections n’ont plus de sens et n cessitent l’introduction d’une nouvelle physique. L’existence d’un scalaire fondamental dans la th orie n’est pas non plus conceptuellement tr s satisfaisante. Pour l’instant, aucun scalaire l mentaire n’a t d tect dans la nature. Toutes les particules de mati re sont des fermions de spin 1/2. L’origine du m canisme de Higgs est inconnue et l’ chelle lectrofaible para t arbitraire. Elle implique des ajustements fins des param tres et c’est ce qui la rend non-naturelle. Pour rem dier tout ceci, la Technicouleur (TC) fut introduite. Cette th orie repose sur l’existence d’une brisure dynamique (comme pour QCD) de l’interaction lectrofaible. Le boson de Higgs est alors un condensat fortement li de deux fermions tr s lourds, les techniquarks. C’est la valeur non-nulle dans le vide de ce condensat qui brise spontan ment la sym trie lectrofaible et r alise le m canisme de Higgs : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (1.24) Ce m canisme de "condensation" s’observe en physique du solide quand un materiau devient supraconducteur et c’est la condensation en paires de Cooper (paires d’ lectrons) qui brise la sym trie $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}}`$. La TC donne donc une origine purement dynamique la brisure et l’ chelle lectrofaible est reli e au couplage technicouleur mobile. C’est une nouvelle interaction, tr s forte, qui se manifeste des nergies de l’ordre du TeV. Sa pr sence r soud le probl me de la divergence quadratique de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$. Dans le mod le minimal, la sym trie de jauge sur laquelle se base la TC est $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ et les nouveaux fermions, U et D, appartiennent la repr sentation fondamentale. Ce sont les analogues des quarks l gers de la sym trie chirale approch e de QCD. Mais la version la plus simple, bas e sur une transposition plus haute nergie de QCD, n’est pas compatible avec les donn es de pr cision lectrofaible. D’autres mod les non-minimaux ont t construits, comme la technicouleur tendue (ETC), la "walking" TC, la "topcolor assisted" TC, mais nous ne les tudierons pas dans ce cours. Le LHC devrait pouvoir d couvrir ou infirmer le principe de la TC et du Higgs composite. #### 1.3.3 Les contraintes sur la masse du Higgs Les donn es de pr cision lectrofaibles sont capables de nous renseigner sur la nature du Higgs et d’une fa on g n rale sur la nouvelle physique. En effet, Veltman a montr la sensibilit des corrections boucles aux particules "invisibles", c’est- -dire trop massives pour tre produites. Par exemple, au niveau d’une boucle, les masses du $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ et du $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$ sont donn es par: $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{sin}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cos}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{sin}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}}{\sqrt{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{G}}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}$$ (1.25) o $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}`$ est la correction radiative calculer. Celle-ci re oit d’importantes contributions du quark lourd $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$: $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{G}}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{8}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\sqrt{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}$$ (1.26) et du boson de Higgs : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\frac{\sqrt{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{G}}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{16}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{11}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{ln}}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (1.27) L’influence de toute particule n c ssaire pour renormaliser le MS, par exemple le quark $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$ et le boson de Higgs, diverge quand sa masse devient infinie. Les tests de pr cision peuvent donc tre sensibles la nouvelle physique m me si les nouvelles particules sont trop lourdes pour tre produites directement. L’influence quadratique du quark $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$ a permis, il y a plusieurs ann es, la pr diction de la masse du top, ce qui a t ensuite confirm e par sa mesure a Fermilab. L’influence du Higgs au niveau d’une boucle n’est que logarithmique, comme nous voyons dans l’ quation (1.27), à cause de ce que l’on appelle l’ crantage de Veltman <sup>37</sup><sup>37</sup>37Au niveau de deux boucles, l’influence de $`\colorbox[rgb]{1,1,1}{$M$}_\colorbox[rgb]{1,1,1}{$H$}`$ devient quadratique aussi.. N anmoins, en utilisant plusieurs donn es diff rentes, on peut tout de m me donner une estimation de la masse du Higgs. En combinant toutes les donn es acquises par le LEP, le Tevatron et SLD nous obtenons la figure (1.3). Les mesures exp rimentales accordent au boson de Higgs une masse : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{114}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{45}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{69}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{G}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}$$ (1.28) Les recherches d’observations directes du LEP donnent une limite inf rieure la masse du Higgs puisque celui-ci n’a pas encore t vu (r gion grise sur la courbe de (1.3)) : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{114}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{G}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{@}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{95}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\%}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (1.29) De plus, nous pouvons aussi utiliser la masse du boson $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ pour contraindre la masse du boson de Higgs. Nous voyons l aussi, sur la figure (1.4), que pour tre consistant avec les mesures de la masse du $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$, la masse pr f r e du boson de Higgs est entre 10 et 200 GeV environ. Les tests de pr cision lectrofaibles donnent des indications claires en faveur d’un boson de Higgs l ger, proche de la limite directe actuelle de 114 GeV. La figure (1.5) donne la gamme de masse du boson du Higgs favoris e par les mesures de diff rentes observables. Nous voyons l aussi que la masse du Higgs est attendue aux environ de la centaine de GeV. Mais il est aussi possible d’ tudier la consistence th orique du MS en fonction de la masse du boson de Higgs car 2 effets s’opposent : les corrections radiatives l’autocouplage $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}`$ du potentiel de Higgs et la stabilit du potentiel. En effet, quand nous prenons en compte les ordres sup rieurs de la th orie des perturbations (les diagrammes boucles), la valeur de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}`$ diverge en fonction de l’ nergie. Il existe donc une nergie seuil partir de laquelle le secteur de Higgs du MS cesse d’avoir un sens, o les corrections deviennent beaucoup plus grandes que la valeur au niveau de l’arbre. De plus, si $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$ est trop faible, l’ volution par le groupe de renormalisation peut rendre $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}`$ n gatif et le potentiel de Higgs est alors instable <sup>38</sup><sup>38</sup>38Le potentiel change de forme et n’est plus limit par le bas.. Les consid rations de stabilit du potentiel et les corrections radiatives contraignent la masse du boson de Higgs dans les deux sens, par le bas et par le haut respectivement. La figure (1.6) montre le "couloir" permis pour la masse du Higgs en fonction de l’ chelle d’ nergie $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}`$. La largeur des bandes correspond l’incertitude th orique sur les r sultats. #### 1.3.4 Sa d tection potentielle au LHC et LC Nous allons maintenant nous tourner vers la d tection du boson de Higgs dans les acc l rateurs de particules. Tout d’abord quelles sont les chances de l’observer au Tevatron, le collisionneur de proton-antiproton du Fermilab a Chicago ? La figure (1.7) montre la sensibilit du Tevatron en fonction de la masse. Un signal sera visible (une vidence $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}`$) si $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$ est inf rieure ou gale 115 GeV mais peut tre pas avant 2007, date laquelle le LHC devrait tre lui aussi op rationnel. L’ valuation de la sensibilit la plus r cente correspond aux lignes les plus fines. Le LHC, avec une nergie disponible dans le centre de masse beaucoup plus importante trouvera de fa on certaine <sup>39</sup><sup>39</sup>39Dans la mesure où, bien sûr, le boson de Higgs existe. le boson de Higgs toutes les masses jusqu’ plusieurs TeV. Apr s plusieurs ann es de mesures, la pr cision attendue sur sa masse est de l’ordre de 1 $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\%}$}`$ et le LHC pourra mesurer beaucoup de ses propri t s ainsi que les param tres li s la brisure de la sym trie lectrofaible. Nous voyons sur la figure (1.8) la sensibilit de l’exp rience ATLAS au LHC pour diff rentes d sint grations du Higgs. Dans le cas d’un collisionneur lin aire $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$, les r sultats seront d’une mani re g n rale compl mentaires. L’ nergie disponible dans le centre de masse est moins lev e mais les bruits de fond sont beaucoup moins importants et donc les incertitudes exp rimentales plus faibles (ce sont des particules fondamentales qui sont acc l r es, pas des protons). Globalement, deux sc narios sont possibles. Soit le Higgs est l ger comme pr vu, soit il est beaucoup plus lourd voire inexistant. Avant de connaitre la r ponse, nous pouvons d j r fl chir aux cons quences de ces deux types de sc nario. ##### Un Higgs l ger? Dans ce cas, un Higgs de l’ordre de 115 GeV demande n cessairement l’introduction d’une nouvelle physique aux alentours de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{6}$}`$ GeV. En effet, le potentiel effectif $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{[}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{]}$}`$ sera "d stabilis " par les corrections radiatives (l’autocouplage du Higgs, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$, finit par changer de signe cause de la renormalisation. Le potentiel n’a alors plus sa forme de "chapeau mexicain" et le vide n’est plus stable). Le MS n’est plus valide partir d’une certaine chelle et il faut introduire d’autres particules pour modifier la renormalisation de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$. L’ volution avec l’ chelle d’ nergie $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}`$ des param tres physiques d pend de leur fonction b ta <sup>40</sup><sup>40</sup>40La fonction b ta est le coefficient d’ volution avec l’ nergie que l’on trouve dans les quations du groupe de renormalisation.. Pour $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$, en ne regardant que l’effet du quark top, nous avons : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{16}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\hspace{0.17em}36}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{27}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{9}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\hspace{0.17em}4}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{9}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\hspace{0.17em}2}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (1.30) avec $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$, la constante de renormalisation une boucle du champ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$ : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{16}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{6}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{9}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\hspace{0.17em}2}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (1.31) Quand nous regardons le signe devant le couplage de Yukawa du quark top, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}`$, il parait clair qu’ajouter de nouveaux fermions ne r soudra pas la stabilit du potentiel, bien au contraire (puisque la fonction b ta sera encore plus "n gative" et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$ diminuera plus rapidement). La nouvelle physique requise doit introduire de nouveaux bosons. Dans le lagrangien nous avons alors l’ajout d’un terme ($`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}`$ est le nouveau boson et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$ le Higgs): $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}$$ (1.32) La figure (1.9) donne des exemples d’ chelle critique $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}`$, chelle partir de laquelle $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$ change de signe et le potentiel devient instable, en fonction de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$. Pour une chelle de nouvelle physique M de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}`$ GeV, l’instabilit du potentiel est repouss e tr s loin. De plus, si nous augmentons la valeur de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ de fa on ce que l’instabilit du potentiel n’intervienne pas avant l’ chelle de Planck, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{19}$}}`$ GeV, un ajustement extr mement fin de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ est n cessaire. Si nous passons de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{70.9}$}`$ GeV 71 GeV, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ passe de n gative $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\hspace{0.17em}5}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\times }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{7}$}`$ GeV un changement radical de comportement juste apr s cette limite et devient m me si grande l’ chelle de Planck que les calculs perturbatifs deviennent incertains, figure (1.10). La supersym trie est la seule th orie connue qui r soud de fa on naturelle, et tous les ordres (toutes les boucles), le probl me li cet ajustement fin. Dans les th ories supersym triques, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$ est uniquement reli e aux couplages de jauge $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\hspace{0.17em}2}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (1.33) L’ volution, montr e sur la figure (1.11), est modifi e par les liens entre les fermions du MS (le top surtout) et les superpartenaires fermioniques et implique une remarquable stabilit de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$, donc du vide lectrofaible. Ainsi, un boson de Higgs l ger, d’une masse d’environ 115 GeV pointerait de mani re tr s forte l’existence de la supersym trie comme nouvelle physique au-del du MS. ##### Un Higgs lourd ou pas de Higgs? Si nous n’observons pas de boson de Higgs avec une masse inf rieure 130 GeV, o nous sommes-nous tromp s? Voici quatres pistes que nous ne d taillerons cependant pas pour viter d’alourdir le cours. Si le boson de Higgs n’a pas la masse attendue, peut- tre n’avons nous pas correctement interpr t les donn es de pr cision lectrofaibles . En effet, les r sultats de l’exp rience NuTeV (collisions in lastiques $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}`$) montrent un l ger d saccord de quelques carts standards par rapport au fit dans le MS. De plus, il y a aussi un d saccord sur les valeurs favoris es de la masse du Higgs entre les asym tries leptoniques et hadroniques de la masse du Z. Si ces d saccords s’av rent confirm s c’est que nous ne comprenons pas tout fait les effets hadroniques ou alors qu’il y a de la nouvelle physique derri re ces anomalies. Et dans ce cas, il n’y a pas de pr diction vraiment solide en faveur d’un boson de Higgs l ger puisque les analyses se basent sur le fit dans le cadre strict du MS. Deuxi me point, le MS est une th orie effective. Nous pouvons donc nous attendre ce que les effets de nouvelle physique se fassent sentir via l’introduction dans le lagrangien de termes non-renormalisables. Nous avons alors un lagrangien de cette forme : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{r}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{o}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}}{\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{n}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{o}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{o}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}}{\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}{\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒪}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (1.34) o $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒪}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}}`$ sont des op rateurs de dimension de masse la puissance ($`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}`$) traduisant les interactions suppl mentaires. Le lagrangien tant de dimension de (masse)<sup>4</sup>, ces op rateurs sont supprim s par une chelle de masse $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ lev e la puissance $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}`$. Le fit global du MS sugg re, pour un Higgs l ger, que les coefficients des nouvelles interactions sont faibles : $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}`$ TeV pour des $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ de l’ordre de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. Mais il reste possible malgr tout qu’une nouvelle physique cette nergie rende $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$ grande. La figure (1.12) nous le montre : il existe un troit couloir pour une valeur lev e de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$ encore compatible avec les donn es actuelles. Les mod les "Little Higgs" reposent sur l’annulation des divergences quadratiques de la masse du Higgs et des divergences dues aux boucles de bosons de jauge par l’introduction de nouvelles particules : un pseudo-quark <sup>41</sup><sup>41</sup>41On l’appelle ainsi car son introduction a pour but de contrer l’effet du quark top. Cependant, il a un spin 1. top T lourd ainsi que d’autres bosons de jauge $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^{^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}`$ et des bosons de Higgs. Le spectre de masse des particules est , en plus d’un boson de Higgs relativement l ger : $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{200}$}`$ GeV, $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{T}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{200}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^{^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{6}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{T}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{200}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{T}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (1.35) Pour obtenir de nouveaux bosons, le MS est inclu dans un groupe de jauge plus large, bris spontan ment. Ces mod les, bien qu’incomplets, resolvent partiellement le probl me de la hi rarchie des masses et constituent des alternatives int ressantes la supersym trie pour des nergies inf rieures 10 TeV environ <sup>42</sup><sup>42</sup>42Au delà de cette chelle d’ nergie, il faudrait enrichir le mod le ”Little Higgs” pour avoir une th orie plus compl te.. De plus, leurs cons quences ph nom nologiques seront observables au LHC ce qui rend l’ tude de ces mod les attractive. La derni re option, plus radicale, est qu’il n’existe pas de boson de Higgs du tout. Des mod les sans Higgs ("Higgsless") ont d j t construits mais pr disent tous une diffusion WW forte l’ chelle du TeV. En effet, le r sultat du calcul des sections efficaces des processus de type $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ (diffusion WW) diverge si il n’y a pas de particule de spin 0 pour annuler les divergences dues aux changes de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$. Dans le cadre quadridimensionel standard, c’est a priori incompatible avec les donn es de haute pr cision. Mais en ajoutant une dimension spatiale suppl mentaire et en postulant des conditions aux bords qui brisent la sym trie lectrofaible <sup>43</sup><sup>43</sup>43Sans rentrer dans le d tail, puisque le boson de Higgs n’existe pas, il nous faut trouver un autre moyen de briser la sym trie lectrofaible. Ceci est fait en ”jouant” sur la g om trie de la dimension suppl mentaire via des conditions aux bords., on d cale la pr diction d’une diffusion forte WW vers 10 TeV. L’observation de cette diffusion est donc un test crucial des mod les Higgsless, qui prouvent toujours quelques difficult s avec les mesures de haute pr cision. ### 1.4 Conclusions et carte de route Le MS est une construction math matique dot e d’un fort pouvoir pr dictif malgr le grand nombre de param tres libres. Dans ce mod le, il reste n anmoins connaitre de fa on certaine la brisure de sym trie lectrofaible. Le responsable pr sum de cette brisure, le boson de Higgs, n’a en effet pas encore t observ exp rimentalement. Mais mis part ceci, le MS est en extraordinaire accord avec les donn es exp rimentales accessibles actuellement ( nergies aux environs de 100 GeV). Nous savons cependant que de la nouvelle physique doit appara tre l’ chelle du TeV et au-del . Le MS ne peut pas donner de r ponses toutes les questions et les myst res de la physique des particules, un nouveau mod le doit prendre le relais. Dans ce contexte le MS est l’approximation basse nergie d’une description plus fondamentale mais incontournable aux nergies sup rieures. Il est alors vu comme une th orie effective. Parmis la liste des questions en suspens que nous avons vu la section 1.1.3, certaines peuvent se regrouper. Globalement, il reste 3 grands probl mes qu’une "th orie du tout" doit r soudre : -Le probl me de l’unification: Pourquoi observons nous 4 interactions et comment pouvons nous les unifier ? Existe t’il un groupe de sym trie qui peut toutes les englober? Les Th ories de Grande Unification (G.U.T en anglais) existent-elles ? -Le probl me de la masse: Qu’elle est l’origine de la masse des particules ? Sont elles dues au boson de Higgs? Pourquoi sont elles si petites devant la masse de Planck ? -Le probl me des saveurs: Pourquoi y-a-t’il autant de quarks et de leptons? Pourquoi y-a-t’il ce m lange observ entre les saveurs? Existe-t’il des sym tries suppl mentaires entre les diff rentes saveurs? De toutes les possibilit s de nouvelle physique tudi es, une en particulier parait incontournable bon nombre de physiciens… ## Chapitre 2 Supersym trie Comme nous venons d’en discuter, le Mod le Standard (MS) est une description valable des ph nom nes physiques des nergies inf rieures quelques centaines de GeV. Au-del , la supersym trie pourrait jouer un r le important et nous nous rapprochons de plus en plus de sa possible d couverte. En effet, le LHC va sonder les nergies de l’ordre du TeV et nous verrons dans ce chapitre que c’est cette nergie que la supersym trie est sens e faire son apparition. Ce cours est principalement orient vers la pr sentation et la discussion des mod les supersym triques. Nous tenterons de discuter clairement de ce que peut apporter la supersym trie en physique des particules. Nous commencerons par introduire la supersym trie de fa on historique et nous donnerons, sans trop de d tails, les principales motivations de son utilisation. Une section plus formelle suivra, o nous pr senterons la structure g n rale d’une th orie physique avec supersym trie. Nous poursuivrons le cours par une partie galement th orique mais appliqu e la physique des particules basse nergie (autour du TeV). Nous y trouverons les diff rents mod les, comme le Mod le Standard Supersym trique Minimal (MSSM), qui servent de base l’analyse de la ph nom nologie. C’est tout naturellement que nous aborderons ensuite l’aspect exp rimental, avec d’abord les premi res contraintes exp rimentales puis les sc nario possibles de la d tection de la supersym trie. ### 2.1 Histoire et motivations #### 2.1.1 Qu’est-ce que la supersym trie Comme son nom l’indique, la supersym trie est tout simplement une nouvelle sym trie et l’effet d’une transformation de supersym trie est de transformer un tat bosonique en un tat fermionique et vice-versa, avec $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ o S est le spin. Si nous appelons Q le g n rateur de la supersym trie qui r alise la transformation, alors tr s sch matiquement $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{o}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{o}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{o}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}`$ (2.1) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{o}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{o}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{o}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}`$ (2.2) #### 2.1.2 L’introduction de la supersym trie Mais ceci n’est que l’aspect "physique des particules" de la supersym trie qui est une sym trie tout fait g n rale. Plus formellement, la supersym trie est en fait la seule et derni re extension possible du groupe de Poincar des sym tries d’espace-temps et c’est tout d’abord ainsi qu’elle fut d couverte. En effet, l’origine, on cherchait combiner les sym tries externes (d’espace-temps, comme les translations) et internes (surtout globales) c’est- -dire tendre le groupe de Poincar par des transformations internes. Il y a eu plusieurs essais dans les ann es 1960 mais en 1967, Coleman et Mandula montr rent de fa on formelle qu’il est impossible de combiner les deux types de sym tries. C’est leur fameux th or me no-go. En fait, il tait sous entendu que c’ tait impossible en utilisant des g n rateurs bosoniques (donc de spin entier) habituels. Mais en 1971, Golfand et Likhtman r ussirent l’extension du groupe de Poincar en utilisant des charges fermioniques, donc de spin demi-entier. C’est la supersym trie. La m me ann e, Neveu, Schwarz puis Ramond propos rent des mod les supersym triques 2 dimensions pour obtenir des cordes supersym triques qui expliqueraient l’origine des baryons. Quelques ann es plus tard, en 1973, Volkov et Akulov appliqu rent la supersym trie aux neutrinos (ils voulaient en faire le fermion de Goldstone) mais il fut montr un peu plus tard que leur th orie, 4 dimensions, du neutrino ne d crivait pas correctement les interactions de basse nergie observ es exp rimentalement. La m me ann e, Wess et Zumino propos rent la premi re th orie des champs supersym trique 4 dimensions de vrai int r t du point de vue ph nom nologique. Puis, ensemble avec Iliopoulos et Ferrara, ils montr rent que la supersym trie permettait de supprimer beaucoup de divergences des th ories des champs usuelles. Ceci a rendu la supersym trie tr s attractive et pendant un temps donn , elle fut beaucoup utilis e pour tenter d’unifier les bosons et les fermions. Par exemple pour unifier les particules de mati re de spin 1/2 avec les particules de jauge de spin 1, ou les particules de mati re et les champs de Higgs, dans les m me supermultiplets. Puis en 1976, deux groupes d couvrirent que la supersym trie locale (la transformation de supersym trie d pend alors des coordonn es dans l’espace-temps) incluait une description de la gravitation. C’est ce que l’on a appell la supergravit . Depuis, la ph nom nologie de la supersym trie a t norm ment tudi e et les th ories bas es sur la supersym trie se sont impos es comme les candidates les plus s rieuses pour la physique au-del du MS. #### 2.1.3 Pourquoi la supersym trie Pourquoi introduire la supersym trie en physique des particules ? Qu’est-ce qui la rend si attractive pour les physiciens des particules ? Les motivations de son introduction en physique des particules sont principalement d’ordre physique avant d’ tre esth tiques. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ En effet, la supersym trie apporte quelque chose de tr s important dans les th ories des champs. Dans le cours pr c dent, nous avons discut des corrections radiatives la masse du Higgs. Quand nous calculons la contribution d’une boucle fermionique comme celle de la figure (2.1 a), nous obtenons <sup>1</sup><sup>1</sup>1Pour le calcul, nous avons pris la forme usuelle du couplage de Yukawa du boson de Higgs scalaire au fermion : $`\colorbox[rgb]{1,1,1}{$y$}_\colorbox[rgb]{1,1,1}{$f$}\colorbox[rgb]{1,1,1}{$H$}\overline{\colorbox[rgb]{1,1,1}{$\psi $}}\colorbox[rgb]{1,1,1}{$\psi $}`$.: $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{16}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{[}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{6}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{ln}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{]}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (2.3) o $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}`$ est un cut-off ultraviolet utilis pour restreindre les impulsions dans la boucle et qui est de l’ordre de l’ chelle de la nouvelle physique au del du MS. Nous voyons que la masse du Higgs diverge et que si nous supposons le MS valable jusqu’ l’ chelle de Planck, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{19}$}}`$ GeV, alors $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}`$ et cette correction est $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{30}$}}`$ fois plus forte que la valeur raisonnable de la masse du Higgs au carr , ($`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ GeV)<sup>2</sup> ! Cette constation est la m me si nous consid rons plut t une boucle d’un champ scalaire S, figure (2.1 b). $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{16}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{[}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{ln}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{]}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (2.4) o $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}`$ est son couplage avec le boson de Higgs. Le probl me qui vient d’ tre nonc est le probl me de la hi rarchie vu au chapitre 1. Que nous apporte la supersym trie dans ce cas ? Si nous regardons de plus pr s les 2 quations pr c dentes, les deux contributions divergentes (leur terme $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$) s’annulent si pour chaque fermion de notre th orie entrant dans la boucle nous avons aussi 2 scalaires avec $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ . Nous verrons juste apr s que c’est exactement ce que la supersym trie se propose d’apporter ! Et de plus, tous les ordres (c’est- -dire quand nous consid rons des corrections plusieurs boucles) ceci est r alis . Il nous reste alors une divergence logarithmique mais qui n’induit pas de probl mes d’ajustements fins. A ce jour, la supersym trie fournit la r solution du probl me de la hi rarchie la plus naturelle et la plus efficace. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ De plus, nous avons vu la fin du chapitre pr c dent que les donn es exp rimentales penchent en faveur d’un Higgs relativement l ger et qu’un Higgs l ger demande, pour stabiliser le vide lectrofaible, que la physique plus haute nergie partage beaucoup de choses en commun avec la supersym trie. Inversement, les calculs faits partir du Mod le Standard Supersym trique Minimal donnent la pr diction $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{130}$}`$ GeV qui est donc compatible avec les ajustements exp rimentaux vus au chapitre pr c dent. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ Les deux premiers arguments ne concernent que le Higgs. Consid rons maintenant les constantes de couplages qui caract risent chaque force fondamentale. Si nous faisons voluer les 3 couplages du Mod le Standard en fonction de l’ nergie, nous observons qu’ils tendent tous se croiser la m me chelle. Enfin presque, c’est quand nous ajoutons la supersym trie dans les calculs d’ volution que les couplages se croisent quasi exactement au m me point (autour de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\times }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{16}$}}`$ GeV). Personne n’est oblig de croire en une "Grande Unification" qui repose sur cette unification des couplages, mais il est tr s tonant d’observer que la supersym trie r alise l’unification aussi pr cis ment. C’est tout de m me un argument fort en faveur de la supersym trie car les th ories de Grande Unification ont beaucoup de caract ristiques et de pr dictions int ressantes pour la physique de basse nergie (elles expliquent par exemple les nombres quantiques des particules vus dans le tableau (1.1) du chapitre pr c dent). $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ Il y a aussi un argument de ce type en faveur de l’existence de la supersym trie dans la nature. Les th ories de cordes, qui fournissent l’heure actuelle une des seules solutions au probl me de la gravitation quantique (avec la gravit quantique boucles), ne peuvent se passer de la supersym trie sans souffrir d’inconsistences physiques et math matiques. L aussi la supersym trie semble un ingr dient essentiel pour construire des th ories coh rentes haute nergie. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ Enfin, la supersym trie peut aussi jouer un r le dans l’explication de la mati re noire. C’est une des seules th ories qui poss dent dans son spectre de particules des candidats s rieux. Ils ont pour nom, neutralino ou gravitino, et la supersym trie pr dit, dans un cadre plausible, leur stabilit . Ainsi, une fois cr s, il ne se d sint grent pas et ne peuvent pas tre observ s directement. Leur nombre peut donc contribuer la masse manquante sous la forme de mati re noire. Cette discussion d montre que la supersym trie est une sym trie tr s s duisante th oriquement, mais surtout tr s utile. Nous n’avons encore aucune preuve de son existence mais au vu de ce qu’elle apporte en physique des particules, il est difficile d’y rester insensible. Nous n’avons donn que des arguments qualitatifs dans cette partie, nous allons maintenant entrer beaucoup plus en d tail. ### 2.2 La structure d’une th orie supersym trique #### 2.2.1 Interlude: "spinorologie" Avant de d marrer une pr sentation plus th orique de la supersym trie nous allons dans ce br ve interlude donner nos notations et conventions sur ce qui concerne les spineurs. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ Un spineur de Weyl d crit une particule de spin 1/2 et de chiralit donn e. C’est un spineur 2 composantes. Nous nous efforcerons le plus souvent de les noter par des lettres minuscules grecques avec un indice. Par exemple $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\xi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$,… o $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{1,2}}$}`$. Un spineur $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}`$ ou $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}`$ est par convention de chiralit gauche, le spineur droit est not $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}^{\dot{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}}}`$ ou $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$. A noter que : $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}_{\dot{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}}}`$ (2.5) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}^{\dot{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}`$ (2.6) et que les matrices $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\epsilon }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\epsilon }$}_{\dot{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}}\dot{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\epsilon }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\epsilon }$}^{\dot{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}}\dot{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ permettent de monter et descendre volont les indices spinoriels $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ Un spineur de Dirac se construit avec 2 spineurs de Weyl et r unit les deux chiralit s d’une particule donn e. C’est donc un spineur 4 composantes. Nous les noterons par des lettres majuscules grecques : $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Psi }}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Phi }}$}`$,… En termes de ses spineurs de Weyl, nous avons : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Psi }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\begin{array}{c}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\begin{array}{c}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\\ \overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\eta }$}}^{\dot{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}}}\end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}$$ (2.7) Les op rateurs de projection $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ permettent de s lectionner l’une ou l’autre chiralit : $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Psi }}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Psi }}$}`$. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ Un spineur conjugu de charge est un spineur auquel l’op rateur de conjugaison de charge a t appliqu . Il d crit la m me particule mais sa charge lectrique est oppos e. $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Psi }}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Psi }}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\begin{array}{c}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\eta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\\ \overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}^{\dot{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}}}\end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}$$ (2.8) o la matrice de conjugaison de charge $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}`$ peut s’ crire : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}$$ (2.9) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ Un spineur de Majorana se construit avec un seul spineur de Weyl mais l’englobe dans une notation 4 composantes. Il est gal son conjugu de charge, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Psi }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Psi }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}`$. $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Psi }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\begin{array}{c}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\begin{array}{c}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\\ \overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}^{\dot{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}}}\end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}$$ (2.10) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ La repr sentation des matrices $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}`$ choisie est la repr sentaion de Weyl dans laquelle : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\begin{array}{cc}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\\ \overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}$$ (2.11) avec $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{𝟏}}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{𝟏}}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ o les $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ sont les matrices de Pauli, et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{diag}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{𝟏}}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathbf{,1}}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Nous avons aussi $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\eta }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}`$ o $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\eta }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{diag}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ est la m trique de Minkowski utilis e pour monter et descendre les indices de Lorentz. #### 2.2.2 L’alg bre et les supermultiplets Comme il a t dit pr c demment, la supersym trie combine des transformations spatio-temporelles du groupe de Poincar et des transformations de sym trie interne (d’un groupe de jauge donn ). La supersym trie introduit des nouveaux g n rateurs $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}`$, fermioniques, donc anticommutant entre eux (ce sont des spineurs): deux op rations de supersym trie ne commutent donc pas entre elles. A priori, rien n’interdit l’introduction de plusieurs g n rateurs, mais dans la version simple de la supersym trie il n’y a qu’un seul nouveau couple de g n rateurs, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}`$ et <sup>2</sup><sup>2</sup>2Car les Q et $`\overline{\colorbox[rgb]{1,1,1}{$Q$}}`$ sont des spineurs donc des objets complexes qui se transforment de fa ons diff rentes sous le groupe de Lorentz. $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}}^{\dot{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}}}`$. C’est la "supersym trie $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒩}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$" et c’est la seule dont nous parlerons. Si la raison est ici p dagogique, dans la section suivante nous donnerons les raisons physiques de ce choix. La connaissance de l’algèbre de la supersym trie (et d’une quelconque sym trie) se résume à la connaissance des relations de commutation de tous ses g n rateurs (son alg bre de Lie). Nous avons, en plus des relations de commutation isues de l’algèbre de Poincar , celles qui font intervenir les g n rateurs $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}`$ et $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}}^{\dot{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}}}`$: $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{[}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{]}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{[}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}}^{\dot{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{]}$}`$ (2.12) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}}_{\dot{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\dot{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}`$ (2.13) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}}^{\dot{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}}^{\dot{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ (2.14) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$ (2.15) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}}_{\dot{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_{\dot{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}}}^{\dot{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}}}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}}_{\dot{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}}}`$ (2.16) Quelle est la signification des $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}`$ ? Tout d’abord, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}`$ est une charge au sens du th or me de Noether c’est- -dire la charge conserv e dans la sym trie. Elle commute donc avec l’Hamiltonien du syst me, cf (2.12). Elle poss de un spin 1/2 et peut alors s’ crire sous la forme d’un spineur de Weyl <sup>3</sup><sup>3</sup>3Elle poss de donc 2 composantes complexes.. On pourrait aussi crire $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}`$ sous la forme d’un spineur de Majorana 4 composantes. De plus, si nous regardons le commutateur (2.13), nous avons sch matiquement $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}`$ c’est- -dire que $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ est une translation d’espace-temps. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}`$ peut donc se voir comme la "racine carr " d’une translation dans l’espace-temps. Nous allons maintenant tudier les repr sentations irr ductibles de cet alg bre (les supermultiplets) et en d tailler le contenu. En effet, nous voulons appliquer la supersym trie la physique des particules, il nous faut donc savoir comment ranger nos particules et quelles seront leurs propri t s de transformation. Dans le groupe de Poincar , il y a 2 l ments invariants de Casimir : l’op rateur de spin $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}`$, avec $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}}`$ le vecteur de Pauli-Lubanski, et l’op rateur de masse $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}`$, o $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}`$ est la quadri-impulsion. Dans un multiplet du groupe de Poincar , les particules ont la m me masse et le m me spin. Mais dans l’alg bre supersym trique, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ n’est plus un invariant de l’alg bre. Nous avons $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{[}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{]}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ (2.17) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{[}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{]}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ (2.18) et donc, dans un supermultiplet les particules ont la m me masse mais des spins diff rents. Nous pouvons tout de m me corriger $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ pour obtenir un nouvel invariant dont les valeurs propres sont sous la forme $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}`$ avec $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{,1}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ le nombre quantique de superspin. Ce nouveau $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ est invariant donc chaque repr sentation irr ductible peut tre caract ris e par un couple $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{[}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{]}$}`$ et le lien entre le spin $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$ est d duit de la relation : $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$. Dans un m me supermultiplet, on aura donc des particules de m me masse et de m me superspin. De plus, une propri t importante est qu’il y a galit dans un supermultiplet entre le nombre de degr s de libert bosoniques et fermioniques : $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}`$. Nous pouvons maintenant construire les diff rentes repr sentations : $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ La représentation fondamentale $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{[}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{,\; 0}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{]}$}`$ est appelée supermultiplet chiral (ou scalaire). La valeur $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ implique $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{,0}}$}`$ donc ce supermultiplet contient 2 champs scalaires r els r unis sous la forme d’un champ scalaire complexe (le sfermion), $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}`$ et un champ fermionique de Weyl (de spin 1/2), $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}`$. Ces champs ont la m me masse. Pour que la supersym trie soit pr serv e dans les boucles o les particules ne sont pas sur leur couche de masse (c’est- -dire $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$) il faut que les degrés de liberté fermioniques et bosoniques soient aussi quilibr s dans ce cas (off-shell). En effet, off-shell un fermion de Weyl poss de 4 degr s de libert de spin au lieu de 2 on-shell. Il faut donc ajouter au contenu de cette repr sentation un autre champ scalaire complexe mais qui ne se propage pas (on dit qu’il est auxiliaire et il n’a pas de terme cin tique). On-shell, nous utiliserons l’ quation du mouvement $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ pour l’ liminer. Le contenu total du supermultiplet chiral est donc $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Psi }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (2.19) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ La seconde repr sentation que nous allons utiliser dans la suite est le supermultiplet vecteur (ou de jauge) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{[}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{,\; 1}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{]}$}`$. Son contenu en champ est obtenu de la m me fa on : un fermion de Weyl <sup>4</sup><sup>4</sup>4O de fa on quivalente un fermion de Majorana. (le jaugino), $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}`$, un boson de jauge (de masse nulle), $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}`$, et comme pour le supermultiplet chiral, un champ scalaire r el auxiliaire, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{D}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}`$. $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Phi }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{D}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (2.20) o $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}`$ est un indice adjoint du groupe de jauge : C’est dans ces deux repr sentations que les particules du MS et leurs superpartenaires seront rang s. Nous allons maintenant construire avec ces deux repr sentations une th orie des champs supersym trique. #### 2.2.3 Th orie des champs supersym trique Avant de discuter de mod les en particulier et surtout de l’extension supersym trique minimale du MS (le MSSM) nous allons d’abord pr senter, sans d monstrations trop lourdes, la structure g n rale d’une th orie des champs avec supersym trie. Pour commencer progressivement, nous allons introduire le mod le de Wess et Zumino sans interaction, juste pour voir comment concr tement se transforment les champs. Puis nous introduirons les interactions et ceci nous conduira la notion nouvelle de superpotentiel. Enfin, nous discuterons des champs de jauge dans une th orie supersym trique. A l’issue de cette section, nous devrions poss der suffisamment de bagage th orique pour comprendre comment le MSSM se pr sente et en tudier les aspects concrets comme les pr dictions exp rimentales. ##### Le lagrangien libre globalement supersym trique L’action la plus simple que l’on peut construire avec le supermultiplet chiral est celle du modèle de Wess-Zumino, sans masse et sans interaction. Dans le cas on-shell (sans $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}`$), qui est suffisant quand il n’y a pas d’interaction, nous avons simplement un fermion $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}`$ et un boson scalaire $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}`$ : $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}{\displaystyle \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{o}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ (2.21) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ (2.22) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{o}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}`$ (2.23) Si on applique une transformation supersym trique globale de param tre $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}`$, fermion de Weyl infinit simal ind pendant des coordonn es d’espace-temps ($`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ ), sur le champ scalaire $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}`$, le r sultat doit tre proportionnel au champ fermionique $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}`$ : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{et}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}}_{\dot{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}}}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}^{\dot{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}}}$$ (2.24) $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}}_{\dot{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}^{\dot{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}$$ (2.25) A noter que le fermion infinit simal $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}`$ a la dimension d’une masse la puissance -$`\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}`$ contrairement un fermion de Weyl usuel qui a la dimension (masse)<sup>3/2</sup> : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{[}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{]}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{[}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{]}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{[}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{]}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (2.26) La th orie physique est invariante si l’action est invariante (principe de moindre action). Mais pour que l’action reste invariante sous une transformation de supersym trie, il faut que la somme $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{o}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}`$ soit nulle une divergence totale pr s, qui ne contribuera pas l’action. En consid rant aussi la dimension des 3 champs dont nous disposons, nous voyons que les champs fermioniques se transforment comme: $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{et}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}^{\dot{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^{\dot{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}$$ (2.27) Mais est-ce que cette transformation correspond bien une transformation de supersym trie? Pour s’en convaincre il suffit de partir, soit d’un fermion $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}`$ soit d’un boson $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}`$, et de leur appliquer 2 fois ces transformations. Nous avons la cha ne suivante : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (2.28) c’est- -dire que dans les deux cas l’action de 2 transformations supersym triques successives est quivalent la d rivation donc l’op rateur impulsion $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}`$ (car $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}`$). Nous retrouvons le r sultat de la section pr c dente, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}`$, et donc nos transformations r alisent bien l’alg bre supersym trique. Dans le cas off-shell, l’action $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}`$ est modifi e par l’ajout d’un terme comportant le champ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}`$: $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}{\displaystyle \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{o}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}`$ (2.29) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}`$ (2.30) et les transformations de supersym trie des champs $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}`$ s’en trouvent modifi es. Pour le champ scalaire $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}`$, sa transformation doit faire intervenir le champ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}`$. Remarquons que la dimension du champ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}`$ est une masse au carr . Avec les lois de transformation suivantes, $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}}^{\dot{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_{\dot{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{et}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}^{\dot{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_{\dot{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (2.31) la variation du terme $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}}`$ donne : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (2.32) Cette variation s’annulent bien on-shell avec l’ quation du mouvement $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$. Pour compenser cette variation dans le cas off-shell, la loi de transformation de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}`$ devient : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{et}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}^{\dot{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^{\dot{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}}^{\dot{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (2.33) Les transformations de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}`$ sont inchang es. Nous pouvons v rifier que $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ sans faire r f rence aux quations du mouvement et donc la supersym trie est aussi r alis e off-shell avec ces lois de transformation. Nous avons vu que le champ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}`$ tait auxiliaire et ne servait qu’ l’utilisation off-shell de la supersym trie. En fait, il a aussi un autre r le. En effet, nous sommes partis dans l’explication de la supersym trie mais il ne faut pas oublier qu’aux chelles d’ nergies accessibles actuellement nous n’avons pas observ de supersym trie. Si la supersym trie existe dans la nature, elle a d forc ment tre bris e un moment ou un autre. Le champ auxiliaire $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}`$, mais aussi le champ auxiliaire $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{D}$}`$, servent briser la supersym trie. Nous verrons cet aspect au dernier chapitre. Apr s cette section formelle, nous oublierons ces champs auxiliaires en utilisant chaque fois les quations du mouvement pour les liminer des quations. ##### Les interactions du multiplet chiral Nous allons maintenant ajouter notre th orie la possibilit de termes d’interaction entre ces deux types de champs qui composent les supermultiplets chiraux. Nous n’allons pas le d montrer mais le terme d’interaction le plus g n ral, invariant sous les transformations de supersym trie et renormalisable, que nous allons ajouter dans le lagrangien libre vu pr c demment, s’ crit sous la forme : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (2.34) La quantit $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}`$ ne doit d pendre que des champs $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}`$, pour assurer que la variation lors de la transformation de supersym trie du premier terme de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}`$ puisse tre compens e par la variation d’un autre terme. Pour la même raison, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}`$ doit tre compl tement sym trique. La seule forme possible pour $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}`$ est alors : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (2.35) o on d finit le superpotentiel $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ que l’on crit sous la forme : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{6}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}$$ (2.36) dans le cadre d’une th orie renormalisable. Qu’est-ce que ce superpotentiel ? Tout d’abord, sa dimension est celle d’une masse au cube. Il fait intervenir la matrice sym trique de masse $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}`$ des fermions <sup>5</sup><sup>5</sup>5La supersym trie assure que c’est aussi la matrice de masse des bosons scalaires associ s. et la matrice totalement sym trique des couplages de Yukawa $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}`$ entre un scalaire et 2 fermions. Il r sume donc toutes les interactions qui ne sont pas de jauge (c’est- -dire avec les bosons de jauge). C’est, de plus, une fonction analytique des champs complexes $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ c’est- -dire qu’il est fonction de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ mais pas du complexe conjugu $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$. Ceci est tr s important pour la suite. De m me, en imposant que $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}`$ soit invariant sous transformation de supersym trie, on d termine la forme du potentiel $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}`$. En pr sence d’interactions, donc d’un superpotentiel non-nul, les quations du mouvement des champs auxiliaires $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ sont : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (2.37) Nous pouvons les utiliser pour crire le lagrangien sans les champs $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}`$, comme dit plus haut. Le potentiel scalaire $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}`$ de la th orie est : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (2.38) qui est automatiquement non-n gatif puisque c’est la somme de carr s. Si nous employons la forme g n rale (2.36) du superpotentiel, nous avons alors le lagrangien g n ral pour un supermultiplet chiral en interaction : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (2.39) ##### Th orie de jauge supersym trique Le MS, qui est la th orie qui nous int resse et que nous voulons "supersym triser", a, outre des champs fermioniques chiraux (les quarks, les leptons), des champs de jauge de spin 1 (bosons $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$, gluons,…). Dans la section d di e l’alg bre supersym trique nous avons vu que le supermultiplet vecteur pouvait accueillir de tels champs de jauge. Voyons donc comment se comporte un tel supermultiplet, sans et avec interaction. Le supermultiplet vecteur contient un boson de jauge $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}`$, de masse nulle, un fermion de Weyl, le jaugino $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}`$, galement de masse nulle, ainsi qu’un champ scalaire r el auxiliaire $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{D}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}`$ qui est l’analogue du champ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}`$ pr c dent. Ce supermultiplet est dans la repr sentation adjointe du groupe de jauge (de constantes de structure $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}}`$). La forme du lagrangien est compl tement d termin e par la condition d’invariance de jauge et la renormalisabilit : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{D}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{D}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{D}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (2.40) o les d riv es covariantes de jauge $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{D}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}`$ sont donn s par : $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}`$ (2.41) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{D}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}`$ (2.42) comme habituellement pour une th orie de jauge. Ce lagrangien est d j supersym trique et les transformations de supersym trie de param tre $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}`$ pour les champs du supermultiplet vecteur sont : $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\sqrt{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}`$ (2.43) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\sqrt{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\sqrt{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{D}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}`$ (2.44) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{D}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}}{\sqrt{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{D}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{D}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}`$ (2.45) Sans interaction avec aucun supermultiplet chiral, l’ quation du mouvement pour le champ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{D}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}`$ est simplement $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{D}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ que nous obtenons directement du Lagrangien (2.40). Il n’a pas de terme cin tique et ne se propage donc pas. Dans le MS, les champs de jauge interagissent avec les fermions chiraux. Dans notre version supersym trique il nous faut donc consid rer les interactions entre le supermultiplet chiral et le supermultiplet vecteur. Comme dans le cas non-supersym trique, les d riv es usuelles $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}`$ des fermions sont maintenant remplacer par des d riv es covariantes de jauge $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{D}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}`$. De plus, le lagrangien doit comporter des termes suppl mentaires qui traduisent les interactions entre supermultiplets chiraux et vecteurs. Les lois de transformation supersym triques du supermultiplet chiral changent pour prendre en compte la variation des nouveaux termes. L’ quation du mouvement pour $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{D}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}`$ est alors (les $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}`$ sont les g n rateurs du groupe de jauge selon lesquels les supermultiplets chiraux se transforment et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}`$ est la constante de couplage) : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{D}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (2.46) et le potentiel scalaire complet est : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{D}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{D}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (2.47) Ce potentiel scalaire est automatiquement non-n gatif et s’av re tr s important pour la brisure de sym trie. Nous parlons de termes $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}`$ et de termes $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{D}$}`$ pour faire r f rence respectivement au premier et deuxi me terme du potentiel. Ce potentiel est compl tement d termin par les couplages de Yukawa (via le terme $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}`$) et par les interactions de jauge (via le terme $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{D}$}`$). ### 2.3 Les mod les supersym triques basse nergie La section pr c dente tait assez abstraite, nous allons tout de suite appliquer les diff rents r sultats obtenus. Le Mod le Standard fonctionne tr s bien, nous l’avons vu. Nous allons donc juste lui offrir une promotion en le "supersym trisant" et en conservant toutes ses caract ristiques. Le mod le minimal que nous pouvons obtenir est le MSSM (Minimal Supersymmetric Standard Model). Nous pr senterons son contenu en particules (la nomenclature des nouvelles particules), nous expliquerons comment la sym trie lectrofaible peut se briser, et nous d crirons la brisure effective de supersym trie (brisure dite "douce"). Nous aborderons juste apr s les pr dictions typiques du MSSM. Enfin, nous parlerons aussi des variantes possibles du MSSM car la nature a pu tr s bien choisir une voie un peu plus complexe que ce mod le minimal. #### 2.3.1 Les mod les $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒩}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ Comme nous l’avons d j dit dans la section 1.2.2, a priori le nombre de g n rateurs supersym triques $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}`$ que nous pouvons introduire peut tre sup rieur ou gal 1 (nous parlerons de supersym trie $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒩}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$). Apr s tout, les th ories supersym tries $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒩}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ poss dent d’avantages de sym tries et de ce fait il se trouve qu’elles ont moins de divergences ce qui les rend tr s int ressantes. En effet, dans le cas $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒩}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ il n’y a qu’un nombre fini de diagrammes qui divergent et dans le cas $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒩}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}`$ il n’y en a plus du tout ! Une th orie supersym trique $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒩}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}`$ est intrins quement finie. Tout naturellement, nous aimerions donc construire un mod le $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒩}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}`$ englobant le MS. Malheureusement, basse nergie (autour du TeV), les mod les $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒩}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ ne sont pas r alistes. Ils ne permettent pas la violation de la parit que nous observons dans les interactions faibles. En effet, un supermultiplet d’une th orie supersym trique $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒩}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ poss de toujours les 2 h licit s oppos es la fois donc particules "gauches" et "droites" si gent dans le m me supermultiplet. Ce qui implique qu’elles ont les m mes interactions (car elles sont dans la m me repr sentation du groupe de jauge). C’est malheureusement une conclusion contraire aux observations exp rimentales qui nous disent par exemple que l’ lectron "gauche" (qui fait partie d’un doublet dans le MS) n’a pas la m me interaction avec les bosons $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ que l’ lectron "droit" (qui est un singulet d’isospin faible nul et qui ne "ressent" pas l’interaction faible). Les mod les $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒩}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ ne peuvent donc pas d crire la physique des particules basse nergie. #### 2.3.2 La zoologie du Mod le Standard Supersym trique Minimal La sous-section pr c dente nous a enseign que le cas minimal $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒩}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ tait aussi le seul cas r aliste basse nergie pour englober le MS. Les supermultipets dont nous disposons sont : $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ le supermultiplet chiral qui comprend un fermion de spin 1/2 et un boson de spin 0, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ le supermultiplet vecteur qui comprend un boson de spin 1 et un fermion de spin 1/2. Pouvons nous ranger toutes nos particules du MS dans ces multiplets ? Autrement dit, pouvons nous associer les quarks et leptons aux bosons W, Z, au photon, etc ? Malheureusement, cela poserait des probl mes pour la conservation des nombres quantiques. En effet, les bosons de jauge et les fermions n’ont pas les m mes propri t s de transformation sous les groupes de jauge donc poss dent des nombres quantiques diff rents. La supersym trie ne modifie pas ces nombres quantiques, on ne peut donc pas associer un boson de jauge un fermion connu ou inversement. Cela poserait aussi des probl mes pour la conservation d’autres nombres comme le nombre leptonique car les bosons de jauge que nous connaissons ont un nombre leptonique nul contrairement aux leptons. Il nous faut donc inventer des (super)partenaires toutes les particules connues ! Le tableau suivant, (2.1), donne chaque particule connue le nom, le spin et l’abr viation de son spartenaire. Avant de passer aux sections suivantes, nous allons formuler plusieurs remarques. Tout d’abord nous avons aussi, outre les nouveaux spartenaires, au moins deux doublets de bosons de Higgs. Pourquoi a-t-il fallu aussi ajouter des bosons de Higgs ? Dans l’ tude des th ories de champs supersym triques nous avons introduit la notion de superpotentiel. Celui-ci r sume toutes les interactions possibles des particules (mais qui ne font pas intervenir les bosons de jauge) donc en particulier les interactions de Yukawa avec les champs de Higgs. Ce superpotentiel ne peut pas tre fonction de champs complexes conjugu s. Or dans le MS, pour donner une masse aux quarks type "up" nous utilisons un terme $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$. Comme en supersym trie ce genre de terme est interdit nous devons utiliser un nouveau champ de Higgs, d’hypercharge $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$, et le coupler simplement (sans conjugaison complexe) : $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}`$. Ce nouveau champ, apr s brisure lectrofaible nous laissera donc d’autres bosons de Higgs dont certains seront charg s (voir plus loin). Ce nouveau doublet de Higgs est aussi n cesaire pour annuler les possibles anomalies. Deuxi mement, nous savons bien que dans le MS un fermion droit subit un traitement diff rent d’un fermion gauche. Ils auront en supersym trie chacun un supermultiplet avec chacun un spartenaire. Par exemple $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$. Ces deux squarks sont bien diff rents et pour les identifier nous laisserons l’indice de chiralit $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}`$ ou $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$ tout en sachant qu’il n’a pas de sens physique pour une particule scalaire (spin 0 donc une seule h licit $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$). Troisi mement, pourquoi avons nous fait le choix d’avoir des spartenaires de spin inf rieur ? A priori nous aurions pu associer tous les fermions du MS des spartenaires de spin 1 et aux bosons de jauge de spin 1 des spartenaires de spin 3/2. Cependant, introduire une particule de spin 1 signifie introduire une nouvelle interaction et implique un mod le non-minimal. De plus, introduire des particules de spin \>1 rend la th orie non-renormalisable <sup>6</sup><sup>6</sup>6En analysant en th orie des champs ce qui rend les diagrammes divergents, on aboutit une condition pour qu’un terme du lagrangien soit renormalisable: $`\colorbox[rgb]{1,1,1}{$\mathrm{\Delta }$}\colorbox[rgb]{1,1,1}{$=$}\colorbox[rgb]{1,1,1}{$4$}\colorbox[rgb]{1,1,1}{$$}\colorbox[rgb]{1,1,1}{$d$}\colorbox[rgb]{1,1,1}{$$}\colorbox[rgb]{1,1,1}{$$}_\colorbox[rgb]{1,1,1}{$i$}\colorbox[rgb]{1,1,1}{$n$}_\colorbox[rgb]{1,1,1}{$i$}\colorbox[rgb]{1,1,1}{$($}\colorbox[rgb]{1,1,1}{$s$}_\colorbox[rgb]{1,1,1}{$i$}\colorbox[rgb]{1,1,1}{$+$}\colorbox[rgb]{1,1,1}{$1$}\colorbox[rgb]{1,1,1}{$)$}\colorbox[rgb]{1,1,1}{$$}\colorbox[rgb]{1,1,1}{$0$}`$ o $`\colorbox[rgb]{1,1,1}{$d$}`$ est le nombre de d riv es, $`\colorbox[rgb]{1,1,1}{$n$}_\colorbox[rgb]{1,1,1}{$i$}`$ est le nombre de champs du type $`\colorbox[rgb]{1,1,1}{$i$}`$ dans le terme d’interaction et $`\colorbox[rgb]{1,1,1}{$s$}_\colorbox[rgb]{1,1,1}{$i$}`$ leur spin. Si le spin est trop lev on tombe in vitablement sur des termes non-renormalisables.. Enfin, les $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}}`$ ne s’observent pas directement. En effet, ils se m langent et donc n’apparaissent exp rimentalement que des combinaisons de ces jauginos et higgsinos : celles-ci ont pour nom les neutralinos et les charginos : $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ Les neutralinos $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{1,2,3,4}}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ <sup>7</sup><sup>7</sup>7Not s aussi dans la litt rature $`\stackrel{\colorbox[rgb]{1,1,1}{$~$}}{\colorbox[rgb]{1,1,1}{$\chi $}}_{\colorbox[rgb]{1,1,1}{$\mathrm{1,2,3,4}$}}^\colorbox[rgb]{1,1,1}{$0$}`$. sont de charge lectrique nulle et m langent en particulier les fermions $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}}`$, $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$, $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ et $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ Les charginos $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{1,2}}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$ <sup>8</sup><sup>8</sup>8Not s aussi souvent $`\stackrel{\colorbox[rgb]{1,1,1}{$~$}}{\colorbox[rgb]{1,1,1}{$\chi $}}_{\colorbox[rgb]{1,1,1}{$\mathrm{1,2}$}}^\colorbox[rgb]{1,1,1}{$\pm $}`$. sont charg s lectriquement et m langent les $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$ et les $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$. Ces m langes sont d s au fait que les jauginos et higgsinos poss dent les m me nombres quantiques et ne sont pas distinguables s par ment. #### 2.3.3 Le mod le Le MSSM est l’extension supersym trique minimale du MS. Les quarks et les leptons sont alors mis dans des superchamps chiraux avec leurs superpartenaires et ces superchamps forment des supermultiplets charg s sous $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}`$ de la m me fa on que les multiplets du MS. Les bosons de jauge sont quant eux plac s avec leurs superpartenaires fermioniques dans des superchamps vecteurs. Le superpotentiel le plus g n ral, mais minimal, du MSSM est alors : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒲}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒴}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒴}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{D}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒴}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (2.48) La notation $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}`$ signifie que les champs du supermultiplet sont des champs conjugu s de charge $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}`$. Les champs conjugu s apparaissent car nous avons choisi de ne travailler qu’avec des champs gauches. Les champs droits s’obtiennent justement par cette op ration de conjugaison. De plus, les indices de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ ont t supprim s pour ne pas alourdir l’expression. Nous avons en fait $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒴}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}}`$, … Les $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒴}$}`$ sont les matrices de Yukawa, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\times }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}`$ dans l’espace des saveurs et sont sans dimensions. Elles donnent les masses des quarks et leptons ainsi que les angles et phases de CKM apr s la brisure lectrofaible. Les deux champs de Higgs, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}`$, ont t introduit pour respecter la condition d’analycit du superpotentiel <sup>9</sup><sup>9</sup>9C’est- -dire l’absence de champs complexes conjugu s $`\colorbox[rgb]{1,1,1}{$\varphi $}^{\colorbox[rgb]{1,1,1}{$$}}`$. et donner une masse aux particules "up" et "down" ainsi que pour la condition d’annulation des anomalies. Mais une fois que nous avons deux superchamps de Higgs, un terme qui couple les deux peut a priori exister. Ce terme cependant (le couplage $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}`$) donne naissance au probl me "$`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}`$" : ph nom nologiquement, il est de l’ordre du TeV alors que dans le MSSM rien ne le force tre aussi bas. Dans un mod le plus fondamental, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}`$ pourrait tre li l’ chelle de brisure de la supersym trie. Une fois que ce superpotentiel a t crit, nous pouvons trouver toutes les interactions possibles (mais non de jauge) entre les particules et crire le lagrangien d’interaction, gr ce la formule (2.34), ainsi que le potentiel effectif de la th orie, formule (2.47). #### 2.3.4 La brisure douce de la supersym trie Il reste cependant introduire dans le mod le la brisure de la supersym trie. Mais le m canisme et l’ chelle r elle de la brisure sont encore inconnues. Ce que nous pouvons faire c’est param triser basse nergie cette brisure. Ceci se fait en ajoutant des termes au lagrangien qui brise explicitement la supersym trie. La forme g n rale de ce lagrangien de brisure $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{o}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}`$ est : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{o}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{6}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}$$ (2.49) Il brise bien la supersym trie car seuls les scalaires ($`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$) et les jauginos ($`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}`$) ont un terme de masse. La brisure, bien qu’explicite <sup>10</sup><sup>10</sup>10Par opposition spontan e., est dite "douce" ("soft" en anglais) car on peut montrer qu’elle n’introduit pas de divergences quadratiques. Tous les mod les de brisure, qu’ils aient leur origine dans les th ories des cordes ou de supergravit (voir chapitre 4), conduisent basse nergie cette forme de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{o}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}`$. Avec les superchamps du MSSM, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{o}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}`$ s’ crit : $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{o}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}`$ $`{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ (2.50) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}`$ $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\overline{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{D}$}}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{D}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\overline{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{D}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\overline{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\overline{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}}}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{D}$}}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{D}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ Les masses $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ des jauginos sont en g n ral complexes, ce qui introduit 6 param tres. Les $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}}}`$,…, sont les matrices de masse des squarks et sleptons, hermitiennes et de taille $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\times }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}`$ dans l’espace des saveurs, ce qui fait 45 param tres inconnus. Les couplages $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{D}$}`$,…, sont des couplages trilin aires, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\times }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}`$ et complexes donc caract ris s par 54 param tres. Enfin, les couplages bilin aires des Higgs introduisent 4 param tres. En tout, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{o}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}`$ contient 109 param tres inconnus ! La supersym trie introduit donc beaucoup de param tres mais en contrepartie fait intervenir tr s peu de principes. Ce nombre de param tres "soft" peut cependant tre diminu en red finissant les champs gr ce des sym tries ou des hypoth ses suppl mentaires. La mesure des param tres de brisure douce permettra de tester les mod les de plus haute nergie. #### 2.3.5 La brisure lectrofaible et les bosons de Higgs supersym triques ##### Le potentiel scalaire Le secteur de Higgs du MSSM contient 2 doublets complexes : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\begin{array}{c}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\begin{array}{c}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (2.51) La brisure lectrofaible est donc un peu plus complexe que dans le cas du Mod le Standard ( 1 seul doublet). Au niveau classique ("arbre"), le potentiel scalaire effectif s’ crit, apr s plusieurs simplifications que nous ne d taillerons pas : $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ (2.52) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{8}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}`$ Les termes proportionnels $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ tirent leur origine des termes $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}`$ et le terme proportionnel aux couplages de jauge $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ des termes $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{D}$}`$. Les autres termes viennent de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{o}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}`$ (en omettant les autres scalaires qui ne joueront aucun r le ici). Une brisure spontan e de la sym trie lectrofaible ne peut exister avec cette forme de potentiel que si le param tre $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}`$ v rifie : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (2.53) et de plus, pour que le potentiel soit limit inf rieurement, il faut : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (2.54) Ces conditions ne sont valables qu’au niveau "arbre". Quand la brisure de la sym trie a lieu, les deux champs $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ d veloppent une v.e.v : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (2.55) qui sont reli es celle du MS par : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (2.56) On d finit aussi le param tre $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$ par : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{,\; 0}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pi }$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (2.57) Au minimum du potentiel, $$\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (2.58) nous avons alors les deux relations : $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cos}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}`$ (2.59) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cot}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cos}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}`$ (2.60) Elles sont tr s importantes car lient une quantit mesurable, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$, des param tres de brisure douce. Nous pouvons aussi noter que la phase de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}`$ n’est pas d termin e. ##### Les bosons de Higgs supersym triques Les deux doublets complexes comptent 8 degr s de libert donc 8 scalaires r els. Le m canisme de Higgs, quand la brisure lectrofaible se r alise, en utilise 3 pour donner une masse aux 2 bosons $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ et au $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$. Il reste alors 5 bosons de Higgs qui restent dans le spectre : $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ 2 sont neutres et pairs sous une transformation de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ un est neutre et impair sous $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ les 2 derniers sont charg s, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}`$. Les masses de ces bosons de Higgs supersym triques sont au niveau de l’arbre : $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}`$ $`{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\sqrt{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cos}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}`$ (2.61) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}`$ $`{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{sin}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}`$ (2.62) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}`$ (2.63) et de plus la masse du $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ est born e sup rieurement par : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cos}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (2.64) En particulier, cette derni re relation nous donne $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$. Or ces relations ne sont valables qu’ l’arbre. Nous avons d j vu que la masse des scalaires de Higgs subissent des corrections radiatives non-n gligeables au niveau d’une boucle. On a pour la masse de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ au carr : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{ln}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{8}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{f}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (2.65) o $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{1,2}}$}}}`$ sont les masses physiques des stops (qui sont des m langes des tats $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$ et $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}`$) et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{f}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ est une fonction que l’on peut retrouver explicitement dans . La correction $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ d pend quartiquement de la masse du top, ce qui la rend importante. En prenant en compte cette correction, la masse du boson de Higgs supersym trique le plus l ger, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$, pour des masses de sparticules autour d’un TeV, est : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{130}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{GeV}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (2.66) Dans la figure (2.3), nous pouvons voir $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}`$ en fonction de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}}`$ pour diff rentes valeurs de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$. Ce boson de Higgs supersym trique le plus l ger a donc toutes les caract ristiques n cessaires pour tre le boson de Higgs favoris par les ajustements des donn es lectrofaibles ! Ceci constitue une pr diction importante de la supersym trie. Il y en a d’autres et nous allons en voir tout de suite un aper u. #### 2.3.6 La $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$-parit et la mati re noire ##### La $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$-parit Purement sur des consid rations d’invariance de jauge, d’invariance de Lorentz et de renormalisabilit , nous pouvons introduire dans le superpotentiel (2.50) d’autres termes qui n’ont pas de correspondance avec le MS et qui ne conservent ni le nombre baryonique, ni le nombre leptonique <sup>11</sup><sup>11</sup>11La conservation de $`\colorbox[rgb]{1,1,1}{$B$}`$ et $`\colorbox[rgb]{1,1,1}{$L$}`$ dans le MS est accidentelle mais n’est a priori pas obligatoire dans ses extensions (supersym triques ou pas). En effet, dans le MS, il n’existe pas de termes renormalisables qui violent $`\colorbox[rgb]{1,1,1}{$L$}`$ ou $`\colorbox[rgb]{1,1,1}{$B$}`$, ce qui n’est forc ment le cas dans d’autres th ories. Cependant, dans le MS, il y a des possibles interactions non renormalisables qui violent $`\colorbox[rgb]{1,1,1}{$L$}`$ ou $`\colorbox[rgb]{1,1,1}{$B$}`$.. Ces termes sont : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒲}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{D}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{D}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{D}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (2.67) o $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}`$ sont des couplages inconnus sans dimensions. Cependant, une combinaison des deuxi me et troisi me termes pourrait conduire une d sint gration rapide du proton, alors que le proton est tr s stable <sup>12</sup><sup>12</sup>12Sa dur e de vie exc de $`\colorbox[rgb]{1,1,1}{$10$}^{\colorbox[rgb]{1,1,1}{$33$}}`$ ans !. Ph nom nologiquement, il faut pouvoir s’assurer que ces termes soient supprim s : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\prime \prime }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒪}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{9}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (2.68) C’est ce qui se passe quand on postule une nouvelle sym trie, la $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$-parit . Elle est d finie comme suit : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}}$$ (2.69) o $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}`$ est le spin. C’est un nombre quantique multiplicatif, les particules du MS sont alors paires et les sparticules impaires. Si cette $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$-parit est conserv e, elle implique plusieurs choses importantes : $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ les sparticules sont produites par paires, par exemple : $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{X}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ une sparticule se d sint gre en une autre sparticule (ou en un nombre impair), par exemple : $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}}`$, $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ la sparticule la plus l g re, la LSP, est stable. ##### La LSP et la mati re noire Dans le MSSM, la $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$-parit est conserv e et la sparticule la plus l g re est stable donc ne peut pas se d sint grer. La LSP a un fort inter t car pourrait constituer la majorit de la mati re noire froide favoris e par les mod les cosmologiques de formation des structures (galaxies,…). En effet, au fur et mesure du refroidissement de l’Univers (son expansion) les particules se sont d sint gr es, ont form des baryons, atomes etc, sauf les neutrinos et la LSP. Cette derni re nous parait "invisible" car elle n’interagit que tr s peu avec la mati re et elle contribue alors la densit relique de mati re noire dont les bornes sont, d’apr s les mesures r centes de WMAP , $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0.094}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Omega }}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0.129}$}`$ ( 2$`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}`$). L’existence d’une LSP est une pr diction tr s importante mais sa contribution totale la densit de mati re noire d pend des param tres du MSSM. Les particules candidates sont le neutralino le plus l ger qui est de spin 1/2 et le gravitino qui est de spin 3/2. Le sneutrino le plus l ger a d j t exclu par les recherches directes au LEP. Toutes sont des particules neutres car si elles taient charg es, elles auraient d tre d tect es par les mesures d’isotopes lourds anormaux car une LSP aurait pu se lier au noyau. En effet, si la LSP est charg e lectriquement ou color e, le nombre de ces isotopes lourds anormaux par rapport au nombre d’isotopes normaux devrait tre sup rieur $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{6}$}`$. Exp rimentalement, ce rapport est inf rieur $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{15}$}}`$ voir m me $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{30}$}}`$ ! Pour l’instant, m me si le neutralino est plus " la mode", la nature de la LSP diff re grandement selon les sc nario et les param tres de la supersym trie. #### 2.3.7 Les variantes du mod le minimal Le MSSM n’est pas le seul mod le qui a t construit et ce n’est bien s r pas le seul pouvoir tre r alis dans la nature. Dans les analyses et tudes de la ph nom nologie au-del du Mod le Standard, on parle souvent de violation de la $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$-parit et du NMSSM. Nous allons les pr senter bri vement dans cette section. ##### La violation de la $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$-parit Dans le cas ou la $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$-parit est viol e, les termes de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒲}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}}`$, eq (2.67), donneront des contributions tous les processus. Par exemple, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}`$, les oscillations $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}}`$ ou la d sint gration rare $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{K}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}`$ (voir figure 2.4). Les param tres $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}`$ sont en tout au nombre de 45 (9+27+9) et les bornes sup rieures sont typiquement de l’ordre de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\times }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{100}$}`$ GeV). Ils sont cependant difficiles contraindre car ils apparaissent sous forme de produits, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}`$ par exemple, et nous ne pouvons les isoler que sous certaines hypoth ses <sup>13</sup><sup>13</sup>13Par exemple, supposer qu’il existe une hi rarchie entre les couplages qui violent $`\colorbox[rgb]{1,1,1}{$B$}`$ ou $`\colorbox[rgb]{1,1,1}{$L$}`$ et une hi rarchie selon les diff rentes g n rations de quarks et leptons. Ceci semble raisonnable car les limites obtenues partir des donn es exp rimentales sugg rent qu’il ne peut y avoir violation de $`\colorbox[rgb]{1,1,1}{$B$}`$ et $`\colorbox[rgb]{1,1,1}{$L$}`$ simultan ment.. La ph nom nologie de la violation et de la conservation de la $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$-parit sont tr s diff rentes et tr s riches, leur recherche est donc d j en cours au Tevatron et aussi pr vue aux futurs collisionneurs. ##### Le NMSSM Le NMSSM ("next-to-minimal-supersymmetric-standard-model") est la plus simple extension du MSSM. Dans ce mod le, seul le contenu en particules est modifi car on ajoute un nouveau supermultiplet chiral $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}`$, singulet de jauge. Le superpotentiel est alors : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒲}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{6}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒲}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (2.70) Le supermultiplet $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}`$ contient la fois un fermion chiral et son partenaire scalaire. Le principal inter t du NMSSM est de proposer une solution au probl me "$`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}`$". En effet, en supposant que la partie scalaire de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}`$ d veloppe une valeur dans le vide non-nulle $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$, le terme dans (2.70) donne un $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}`$ effectif : $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$. Or $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}`$ apparait aussi dans $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{o}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}`$ et sa v.e.v est naturellement de l’ordre de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{o}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒪}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ TeV, la masse typique des scalaires et jauginos. Ainsi, la valeur effective de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}`$ est de l’ordre de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{o}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}`$ plut t que d’ tre un param tre libre et ind pendant de la brisure de supersym trie. Ph nom nologiquement le NMSSM diff re du MSSM parce qu’il permet un boson de Higgs l ger plus lourd. De plus, le scalaire $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}`$ peut a priori se m langer avec les autres scalaires du MSSM et former 5 tats de neutralinos. Ainsi, les signatures exp rimentales des sparticules peuvent changer radicalement. ### 2.4 Les premi res contraintes exp rimentales Comme nous avons vu, le secteur de la brisure douce du MSSM compte une centaine de param tres. L’analyse des donn es et l’extraction de valeurs exp rimentales pour ces param tres est alors difficile. Une hypoth se simplificatrice, la base du mod le CMSSM (constrained-minimal-supersymmetric-standard-model), est de supposer l’universalit une certaine chelle : $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ des masses de tous les jauginos : $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ des masses des scalaires : $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{𝟏}}$}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ et que les couplages trilin aires soient reli s par un param tre universel $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ : $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}`$. Ainsi, le passage du MSSM au CMSSM fait passer de plus d’une centaine de param tres seulement 5 param tres ! Cette hypoth se est tr s pratique d’un point de vue ph nom nologique, bien que discutable d’un point de vue purement th orique. Le CMSSM et la simplification de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{o}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}`$ sont en fait inspir s des mod les de supergravit o la brisure de supersym trie se fait par la m diation de la gravit (chapitre 4). Pour ces raisons nous nous placerons dans ce mod le jusqu’ la fin du chapitre pour discuter des contraintes exp rimentales. #### 2.4.1 Les acc l rateurs Les premi res contraintes exp rimentales sur les mod les supersym triques viennent du fait qu’aucune sparticule n’a t d couverte directement, ni au LEP ni au Tevatron. Ceci implique que les masses des superpartenaires soient sup rieures une centaine de GeV environ, selon les cas. Les contraintes sur les sparticules sont aussi obtenues via la limite exp rimentale sur la masse du Higgs. $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{114}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{G}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (2.71) En effet, la masse du Higgs est tr s sensible la masse des particules (celle du stop et du top surtout) circulant dans les boucles, comme nous avons vu pr c demment. Ces deux types de contraintes sont r sum es sur la figure (2.5). La limite sur $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$ donne une limite sur la masse des jauginos $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}`$ qui entre dans le calcul de la masse du stop. De plus, les d sint grations $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}`$ ont un taux de branchement en accord avec le MS. Les boucles de sparticules (de charginos et de Higgs charg s principalement) ne doivent donc contribuer que tr s peu, ce qui contraint aussi leurs masses. Cette contrainte venant de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}`$ est aussi trac e sur la figure (2.5). Enfin, le moment magn tique anormal du muon ($`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$) peut tre utilis comme autre r sultat exp rimental contraignant l’espace des param tres, et cette contrainte a t report e dans la figure (2.5). Mais l’observation d’une r elle d viation de ($`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$) par rapport la pr diction du MS est sujette d bats actuellement dans la communaut des physiciens, la contrainte pr c dente est donc prendre avec prudence. #### 2.4.2 La cosmologie La cosmologie joue un r le de plus en plus important en physique des particules. Quand la $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$-parit est conserv e, la supersym trie apporte un candidat, la LSP, pour la mati re noire froide. Pour respecter les mesures cosmologiques de WMAP sur la densit totale de mati re noire $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Omega }}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$, la contribution de la LSP $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Omega }}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ ne peut pas exc der 0.129 . Les figures (2.5) incluent aussi les r gions permises de la densit de mati re noire d’origine supersym trique par WMAP. De plus, la LSP ne doit pas tre charg e, et cette condition est aussi prise en compte, la r gion triangulaire fonc e est la r gion exclue. ### 2.5 La d tection de la supersym trie aux collisionneurs #### 2.5.1 Les benchmarks Pour tudier les possibilit s de d couverte de la supersym trie aux futurs collisionneurs, il est tr s utile de faire appel des "benchmarks", c’est- -dire des mod les tests dont les param tres sont fix s. Ceux ci servent de rep res et permettent de focaliser la discussion. Ces points de l’espace des param tres doivent tre bien choisis de mani re illustrer les diff rentes possibilit s. Les r gions permises par WMAP et les autres mesures tracent des sortes de bandes fines dans l’espace des param tres du CMSSM. La supersym trie devrait donc se trouver sur un point de l’une de ces bandes. Une dizaine de points significatifs ont t choisi, nomm s par une lettre de l’alphabet (figure (2.6)). Certains points correspondent de petites masses de sparticules, ce qui est donc favorable une d tection directe, d’autres sont parpill s sur les lignes de co-annihilation et 2 points ont t choisis dans des r gions o les LSP peuvent s’annihiler rapidement (points K et M). 2 points, les "points focus" (E et F) sont des points tr s sensibles la masse du top. #### 2.5.2 Les perspectives au LHC et au LC Bien entendu, la possibilit de d couverte d pend des caract ristiques du collisionneur. Sachant que les masses des sparticules peuvent aller de plusieurs centaines de GeV quelques TeV, le Tevatron n’a qu’une mince chance de d couvrir la supersym trie. En revanche, le LHC et le LC ont toutes les chances de la d couvrir car leurs caract ristiques permettent de couvrir une large r gion de l’espace des param tres. Les r sultats simul s des benchmarks sont donn s sur la figure (2.7) qui indique le type de particules que nous pouvons esp rer observer avec le LHC, le LC et la combinaison LHC+LC ou un collisionneur lin aire plus puissant, le CLIC. Au LHC, la supersym trie semble donc "facile" voir dans la plupart des benchmarks. Dans ce cas, le nombre important de d sint grations en cascade permettra d’observer plusieurs sparticules la fois. Les mesures d’ tats finaux supersym triques permettront d’obtenir les masses des sparticules visibles avec une pr cision d’environ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\%}$}`$. De plus, l’observation d’un grand nombre d’ v nements avec beaucoup d’ nergie transverse manquante pourra indiquer la pr sence de supersym trie avec conservation de la $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$-parit . Toutes ces mesures du spectre donneront acc s aux param tres fondamentaux des mod les comme le CMSSM qui pourront tre utilis s pour calculer certaines des observables de basse nergie. ##### Exemple d’utilisation des r sultats Dans les cas les plus favorables, on pourra estimer les param tres du MSSM au LHC (et peut- tre aussi avec un collisionneur lin aire ?) et calculer par exemple la densit de mati re noire. Juste pour donner un exemple, le spectre de masse suivant, correspondant au benchmark "B", pourra tre mesurable au LHC avec les pr cisions indiqu es (les masses sont en GeV) : $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{595.1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{8.0}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{540.3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{8.8}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{520.4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{11.8}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{491.9}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{7.5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{524.5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{7.9}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{202.3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5.0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{143.1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4.8}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{132.5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{6.3}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{96.2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4.8}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{176.9}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4.7}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{377.9}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5.1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}`$ (2.72) En supposant le signe de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}`$ connu, en posant <sup>14</sup><sup>14</sup>14$`\colorbox[rgb]{1,1,1}{$A$}_\colorbox[rgb]{1,1,1}{$0$}`$ n’a de toute fa on que peu d’impact sur $`\colorbox[rgb]{1,1,1}{$\mathrm{\Omega }$}_{\colorbox[rgb]{1,1,1}{$L$}\colorbox[rgb]{1,1,1}{$S$}\colorbox[rgb]{1,1,1}{$P$}}\colorbox[rgb]{1,1,1}{$h$}^\colorbox[rgb]{1,1,1}{$2$}`$. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ on trouve : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{103}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{8}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{GeV}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{240}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{GeV}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10.8}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (2.73) Ce qui donne (voir figure 2.8) : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Omega }}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0.11}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0.03}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0.02}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (2.74) Dans le calcul de la densit de LSP, la limite exp rimentale de WMAP ne sera pas forc ment satur e. Il pourra toujours y avoir d’autres contributions non n gligeables $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Omega }}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ qu’il nous faudra trouver. De plus, le CMSSM n’est pas forc ment le mod le le plus r aliste. Cependant, c’est un des mod les les plus simples pour tudier la physique au-del du MS mesurable dans les collisionneurs venir. ### 2.6 Conclusions La supersym trie est pour l’instant hypoth tique. Seulement, beaucoup d’arguments sont en faveur de son existence. Ces arguments sont essentiellement th oriques mais la supersym trie est aussi sugg r e exp rimentalement. D’un c t elle r sout le probl me de la hi rarchie des chelles, et est essentielle pour les th ories des cordes. De l’autre, la supersym trie ajout e au MS r alise partir des donn es du LEP l’unification des couplages de jauge haute nergie. Elle favorise aussi un boson de Higgs l ger qui est pr f r exp rimentalement. De plus, sa ph nom nologie est tr s riche : elle introduit une foule de nouvelles particules avec lesquelles de nouvelles interactions sont possibles. Elle contribue aux processus rares violant la saveur leptonique et la sym trie $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}`$. Le concept de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$-parit induit aussi une riche ph nom nologie comme par exemple une candidate pour la mati re noire (la LSP). Tout ceci l’a rendue extr mement populaire. Les consid rations de stabilit de la masse du Higgs envers les corrections radiatives sugg rent des masses de l’ordre du TeV pour les spartenaires. Ainsi, elle est tr s attendue au LHC ! La supersym trie s duit car elle permet d’unifier les bosons et les fermions. C’est aussi la derni re sym trie de l’espace-temps 4D ne pas tre (encore) observ e dans la nature, et elle peut inclure naturellement la gravitation (th ories de supergravit , voir chapite 4). Cependant, il est vrai qu’elle introduit beaucoup de nouvelles particules, dont des scalaires et ceux-ci n’ont encore pas t observ s dans la nature. De plus, la supersym trie seule n’explique pas tout : les masses, et la physique des saveurs d’une mani re g n rale, ne sont pas expliqu es. Peut tre faudrait-il ajouter une Grande Unification, des dimensions suppl mentaires ? Il reste encore beaucoup de choses tudier dans la supersym trie, et notament sa brisure. Exp rimentalement aussi, la d couverte de la supersym trie, du mod le (MSSM, NMSSM,…) et de tous ses param tres est un d fi pour les ann es venir. Dans quelques ann es nous serons enfin capable de tester l’id e de la $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$-parit , de la LSP, nous pourrons voir les connexions entre la supersym trie et la cosmologie mais surtout nous devrions enfin conna tre le m canisme et l’origine de la brisure lectrofaible. ## Chapitre 3 Physique des neutrinos et al. Nous avons d j discut des neutrinos et de leur impact sur le Mod le Standard. La physique des neutrinos est riche et les cons quences du m lange des neutrinos sont nombreuses. Ce chapitre aborde donc des sujets tr s divers mais tous li s : les oscillations de neutrinos, la violation de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}`$ leptonique, la violation des nombres leptoniques, la leptogen se et l’inflation cosmologique. Nous verrons ainsi l’importance des sp cificit s du secteur des neutrinos sur la physique des particules et la cosmologie. ### 3.1 Les Masses et oscillations des neutrinos #### 3.1.1 L’ chelle de masse des neutrinos Dans le Mod le Standard, nous connaissons maintenant plus ou moins pr cisemment les masses de tous les quarks, de tous les leptons charg s mais pas du tout celles des neutrinos. A vrai dire, on a m me longtemps cru qu’ils n’avaient pas de masse. De plus, jusqu’ pr sent nous n’avons que des limites sup rieures sur les masses des 3 neutrinos. De la d sint gration $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$ du Tritium nous savons que : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2.5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{eV}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (3.1) L’exp rience KATRIN devrait pouvoir sonder des masses de l’ordre de 0.5 eV. Ensuite, des mesures de la d sint gration $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}`$, nous avons : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{190}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{KeV}}$}$$ (3.2) et il y a des projets pour baisser cette limite d’un facteur $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{20}$}`$. Finalement, des mesures de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}`$, nous avons : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{18.2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{MeV}}$}$$ (3.3) qui devrait dans le futur tre am lior e jusqu’ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$5 MeV. Les masses des neutrinos sont donc plus faibles que celles de la plupart des fermions du MS. Mais les limites astrophysiques sont beaucoup plus contraignantes que ces limites obtenues en laboratoire. Les donn es extraites de l’observation des structures grande chelle dans l’univers (galaxies, amas de galaxies,…) peuvent tre utilis es pour obtenir une limite sup rieure de 1.8 eV sur la somme des masses des neutrinos. Cette limite a t r cemment am lior e par WMAP en : $$\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0.7}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{eV}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (3.4) Dans le cas o la limite est satur e et o les neutrinos sont d g n r s en masse, cel donne une masse individuelle d’environ 0.23 eV, 2000 fois plus petite que celle de l’ lectron ! M me si l’extraction de la masse des neutrinos partir d’observations astronomiques d pend du mod le et de notre connaissance actuelle de la cosmologie, cette limite est robuste. Une autre mani re int ressante d’obtenir des limites sur les masses des neutrinos vient de la recherche de la d sint gration double-$`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$ sans neutrinos ($`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}`$). Sa mesure contraint la somme des masses de Majorana des neutrinos gauches, pond r es par leurs couplages l’ lectron : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0.35}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{eV}}$}$$ (3.5) Les futures exp riences ont l’intention d’am liorer cette limite jusqu’ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0.01}$}`$ eV. Sur la figure (3.1) nous pouvons voir la corr lation entre $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Omega }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}}`$, les donn es des oscillations et de WMAP permettent aussi de contraindre la somme des masses de Majorana des neutrinos gauches. On a $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0.001}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Omega }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0.07}$}`$ donc on s’attend d’apr s la figure avoir peu pr s 0.01 eV $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ 1 eV. #### 3.1.2 Les oscillations des neutrinos ##### La Matrice MNS Nous avons voqu au chapitre 1 le ph nom ne des oscillations entre les saveurs de neutrinos. Celui-ci vient de la diff rence entre les tats propres de masses de ceux de saveurs. Il existe une matrice unitaire qui relie ces deux types d’ tats, la matrice MNS, telle que : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{1,2,3}}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒰}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (3.6) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒰}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}`$ est la matrice de rotation dans l’espace des saveurs $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}`$ (voir figure (3.2)) et peut se param triser en fonction de 3 angles et 3 phases : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒰}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\begin{array}{ccc}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{23}$}}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{23}$}}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{23}$}}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{23}$}}\end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\begin{array}{ccc}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{13}$}}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{13}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{13}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{13}$}}\end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\begin{array}{ccc}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{12}$}}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{12}$}}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{12}$}}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{12}$}}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\begin{array}{ccc}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}$$ (3.7) o $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cos}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{sin}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}`$. La premi re matrice est mesurable par les exp riences sur les neutrinos atmosph riques <sup>1</sup><sup>1</sup>1C’est- -dire des neutrinos $`\colorbox[rgb]{1,1,1}{$\nu $}_\colorbox[rgb]{1,1,1}{$\mu $}`$ produits dans l’atmosp re terrestre par les rayons cosmiques puis d tect s sur terre. comme par exemple SuperKamiokande (SK). L’angle de m lange de ce secteur est à peu près maximal, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{23}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{45}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{23}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\times }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}`$ eV<sup>2</sup> . La seconde matrice est encore inconnue exp rimentalement. Elle est accessible depuis les exp riences bas es sur la production de neutrinos par des r acteurs nucl aires et par des acc l rateurs. Jusqu’ pr sent, les exp riences comme Chooz n’ont tabli que des limites sup rieures sur $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{13}$}}`$ et il n’y aucune information sur la phase $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}`$ de violation de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}`$ (CPV). $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{13}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{15}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ . La troisi me matrice est celle du secteur des neutrinos solaires <sup>2</sup><sup>2</sup>2Ils sont appel s ainsi car ces neutrinos sont produits dans les r actions nucl aires du coeur du Soleil et nous parviennent sur Terre.. Les valeurs favoris es par SK et SNO sont $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{12}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{32}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{12}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{7}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\times }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}`$ eV<sup>2</sup> . Enfin, la derni re matrice contient les 2 phases de Majorana possibles si les neutrinos gauches sont de type Majorana (c’est- -dire gaux aux antineutrinos). Elles sont, en principe, observables dans les exp riences sur la d sint gration $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}`$ mais pas dans les oscillations de neutrinos. D’autres quantit s observables peuvent tre sensibles indirectement ces phases, nous en discuterons un peu plus tard. Si, pour avoir une id e de se qui se passe physiquement, nous prenons le cas simple 2 saveurs, le calcul de la propagation de ces neutrinos dans le vide et de la probabilit de transition donnent : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{sin}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{sin}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1.27}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (3.8) A des longueurs L diff rentes (ou des temps diff rents), les probabilit s d’observer un neutrino de saveur $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$ sont diff rentes. La probabilit oscille en fonction de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ et de l’angle de m lange $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}`$ entre les 2 saveurs ( $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}`$ fix ). Les exp riences bas es sur les oscillations entre saveurs des neutrinos n’ont donc acc s qu’aux diff rences de masses-carr es ainsi qu’aux angles de m langes de la matrice MNS (et la phase $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}`$). ##### Les r sultats exp rimentaux et le bilan Avant SK et KamLAND, seuls des d ficits dans les flux de neutrinos venant des sources astrophysiques avaient t observ s. Mais depuis, des oscillations ont t vues dans les graphes. De plus, deux types d’exp riences, SK/SNO et KamLAND, ont d montr que les neutrinos solaires se comportent de la m me fa on que les neutrinos issus des r acteurs, et SK et K2K ont d montr que les neutrinos atmosph riques se comportent de la m me fa on que les neutrinos issus d’un acc l rateur. M me si le signal n’est pas suffisamment clair pour parler de d couverte ou de confirmation des oscillations de neutrinos, les autres explications qui pr disent des d ficits mais pas d’oscillations sont plus de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}`$ du meilleur ajustement. L’hypoth se des oscillations semble donc tr s solide et vraisemblable. Les exp riences MINOS et OPERA devraient faire beaucoup mieux et dans peu de temps nous pourrons avoir une confirmation nette des oscillations. A l’heure actuelle, nous pouvons dire tre arriv au "consensus" suivant sur notre connaissance des neutrinos : $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ il y a 3 neutrinos l gers, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ leurs masses propres sont rang es dans l’ordre hi rarchique (cf figure (3.3)), $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ le m lange est presque bimaximal : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒰}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\begin{array}{ccc}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\sqrt{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\sqrt{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\\ \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}& \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\sqrt{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}}\\ \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}& \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}& \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\sqrt{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}}\end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}$$ (3.9) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ les phases $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{1,2}}$}}`$ sont compl tement inconnues, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ la nature de Dirac ou de Majorana des neutrinos est aussi inconnue, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ les moments dipolaires des neutrinos sont sans doute faibles, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ les temps de vie des neutrinos sont sans doute longs. Cependant, nous pouvons nous poser la question de l’existence d’autres neutrinos l gers mais st riles <sup>3</sup><sup>3</sup>3L’exp rience LSND aurait vu un troisi me $`\colorbox[rgb]{1,1,1}{$\mathrm{\Delta }$}\colorbox[rgb]{1,1,1}{$m$}^\colorbox[rgb]{1,1,1}{$2$}`$ compl tement diff rent des deux autres, ce qui s’expliquerait par l’existence d’au moins un autre neutrino qui n’aurait pas t d tect avant. (sans interactions faibles), ce qui est confirmer par l’exp rience MiniBoone. Les neutrinos peuvent aussi avoir une hi rarchie invers e, rien n’est encore s r de ce c t . De plus, du c t des angles de la matrice MNS, il reste conna tre $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{13}$}}`$. J-PARC, Double-Chooz et d’autres projets utilisant des r acteurs nucl aires devraient dans un futur proche pouvoir diminuer la limite actuelle. Ensuite, l’angle "solaire" $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{12}$}}`$ n’est pas vraiment maximal, il reste savoir pourquoi. Finalement, la violation de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}`$ dans le secteur leptonique reste d terminer. ##### La violation de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}`$ leptonique Avant de passer aux masses des neutrinos, nous allons encore dire quelques mots sur la violation de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}`$ leptonique. Pour pouvoir d terminer exp rimentalement l’angle $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}`$, il est possible d’utiliser l’observable suivante : $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒫}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒫}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{16}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{12}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{12}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{13}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{12}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{23}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{23}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{sin}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\times }$}`$ (3.10) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{sin}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{12}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{sin}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{13}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{sin}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{23}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}`$ c’est- -dire la diff rence entre les probabilit s d’oscillations de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}`$ et $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}`$. Mais l’extraction de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{sin}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}`$ ne sera possible que si $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{13}$}}`$ est suffisamment grand. Si $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{13}$}}`$ est vraiment faible, l’extraction de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{sin}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}`$ risque d’ tre tr s difficile obtenir et la violation de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}`$ tablir. En revanche, la connaissance de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}`$ pourra entre autres nous donner des indications indirectes pour la leptogen se, un des possibles m canismes responsables de l’asym trie baryonique <sup>4</sup><sup>4</sup>4C’est- -dire le fait que le nombre de baryons soit si grand devant celui des antibaryons, cf la fin du chapitre.. En g n ral, le maximum de l’asym trie (3.10) est quelques centaines ou milliers de kilom tres de la source, ce qui n cessite de tr s longue installations ("Long baseline"). De plus, pour obtenir une mesure claire et sans ambig it s possibles dans l’interpr tation des r sultats, il faudra combiner les informations pour diff rentes distances comme nous le montre la figure (3.4). Avec quels types d’exp riences pouvons nous obtenir les mesures de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{13}$}}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}`$ ? Entre autre, et sans rentrer dans les d tails, avec des usines neutrinos mais aussi avec des beta-beams (faisceaux b ta). Le premier projet est bas sur le stockage de muons qui se d sint grent en $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}`$ et $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}`$. Dans le second, les faisceaux de neutrinos ou d’anti-neutrinos sont cr s par un stockage de noyaux instables qui se d sint grent en cr ant un neutrino ou anti-neutrino dans l’ tat final. Par exemple, $`{}_{}{}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{6}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}`$ et $`{}_{}{}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{18}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}`$. Dans ce cas, le faisceau cr est tr s pur. Une simulation des limites possibles sur $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{12}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}`$ en fonction de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{13}$}}`$ est donn e dans la figure (3.5). En particulier, l’angle $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{13}$}}`$ pourra tre mesur si $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{sin}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{13}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}`$. #### 3.1.3 Les masses des neutrinos ##### Au-del du Mod le Standard Les neutrinos ont une masse, mais dans le cadre strict du MS ceci n’est pas possible. Cependant, puisque le MS n’est qu’une description effective basse nergie de la physique, nous pouvons tout fait ajouter les termes non-renormalisables qui g n rent des masses aux neutrinos, sans m me ajouter d’autres champs. Par exemple, un terme non-renormalisable de la forme <sup>5</sup><sup>5</sup>5C’est le terme non-renormalisable le plus simple. Il en existe bien s r d’autres mais ils font intervenir des puissances de $`\colorbox[rgb]{1,1,1}{$1$}\colorbox[rgb]{1,1,1}{$/$}\colorbox[rgb]{1,1,1}{$M$}`$ plus grandes et des op rateurs plus complexes. : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}}$$ (3.11) o M est une chelle d’ nergie caract ristique de la nouvelle physique responsable de ce terme. Ce terme, la brisure lectrofaible, va g n rer un terme de masse aux neutrinos : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}{\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}{\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}$$ (3.12) Or, une nouvelle interaction comme (3.11) semble non naturelle et peu fondamentale, il faut comprendre l’origine du terme et celle de l’ chelle de masse $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}`$. Num riquement, en prenant $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\sqrt{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{23}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5.10}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ eV et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{200}$}`$ GeV, nous trouvons que $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{15}$}}`$ GeV. L’ chelle de masse de la physique qui fourni le terme effectif (3.11) est donc tr s haute, comparable l’ chelle de Grande Unification. ##### Le m canisme de Seesaw Pour expliquer la masse faible des neutrinos partir d’une haute chelle de masse, nous pouvons utiliser le m canisme de Seesaw (m canisme de la balan oire). C’est un m canisme minimal, c’est- -dire sans nouvelles interactions de jauge, renormalisable, et qui requiert seulement l’introduction de neutrinos lourds singulets de l’interaction faible, donc droits. Ils sont not s $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$. Ces nouveaux neutrinos droits vont pouvoir se coupler avec les neutrinos gauches via un couplage de Yukawa et conduire un terme de masse de Dirac la brisure lectrofaible. Cette masse de Dirac $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{D}$}`$, puisque intimement li e la brisure lectrofaible, est donc naturellement de l’ordre de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$. De plus, les neutrinos droits peuvent eux-m me former un terme de masse de Majorana, de masse $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}`$ qui peut tre beaucoup plus lev e que $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$. Par exemple, elle peut tre de l’ordre de l’ chelle de la Grande Unification, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒪}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. En effet, ce terme de masse n’a rien voir avec la brisure lectrofaible donc ne n cessite pas d’y tre reli . Si nous combinons ces deux types de masse, nous obtenons la matrice de masse dite de Seesaw : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{w}$}}}{\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\begin{array}{cc}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{D}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{D}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{T}}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}\end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\begin{array}{c}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (3.13) ce qui nous donne, en diagonalisant, la matrice de masse effective des neutrinos gauches : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{D}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{D}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{T}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (3.14) o $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{D}$}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}`$ sont des matrices $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\times }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}`$ dans l’espace des saveurs. Afin de retrouver avec 3 masses de neutrinos non-nulles les 2 diff rences de masses-carr es observ es, nous avons besoin d’introduire au moins 3 neutrinos droits $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$. M me si dans ce mod le nous n’avons pas besoin d’une Grande Unification, 3 neutrinos droits apparaissent naturellement dans certains mod les ($`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{O}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ en particulier). En r sum , nous avons donc des masses effectives de neutrinos gauches faibles et inversement proportionnelles aux masses de neutrinos singulets lourds. Le terme (3.11) s’obtient par change de ces neutrinos lourds. Il existe d’autres m canismes de Seesaw, non-minimaux, dans lesquels sont ajout s des champs suppl mentaires. Le m canisme pr c dent est souvent appel le m canisme de Seesaw de type I et n’est pas la seule possibilit d’obtenir un terme effectif basse nergie comme (3.11). Dans le m canisme de Seesaw de type II, on ajoute des Higgs lourds triplets de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}`$ qui sont chang s. Dans le double Seesaw, ce sont d’autres singulets de neutrinos (en plus des 3 du m canisme type I) qu’on ajoute, mais qui ne se couplent pas aux neutrinos gauches. Le m canisme se produit gr ce au couplage avec les neutrinos droits. Ces m canismes sont un peu complexes mais reposent cependant sur le m me principe. Nous continuerons donc l’expos dans le cadre du mod le seesaw de type I. ##### Les alternatives Le m canisme de Seesaw n’est pas le seul pouvoir fournir des masses tr s faibles aux neutrinos, il existe d’autres alternatives, par exemple dans le cas d’une th orie supersym trique avec violation de la $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$-parit . En effet, comme le nombre leptonique n’est pas conserv , il n’est pas tonnant que la supersym trie sans $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$-parit puisse apporter une explication aux masses et m langes des neutrinos. Il y a plusieurs sources possibles aux masses des neutrinos. Au niveau de l’arbre nous pouvons avoir un terme de masse de la forme : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}}}$$ (3.15) o $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ est le couplage bilin aire entre les champs leptoniques et le champ de Higgs de type "up". Au niveau d’une boucle il y a beaucoup de possibilit s, par exemple une boucle de slepton-lepton qui donnerait : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{8}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (3.16) ou bien une boucle de quark-squark, ou encore une boucle de sneutrino-neutralino. L’importance relative de ces contributions d pend du mod le. Nous pouvons facilement tablir une hi rarchie entre les masses, par exemple en g n rant une masse l’arbre gr ce au terme (3.15) et les deux autres par des termes du genre (3.16). Nous n’irons cependant pas plus loin car c’est le m canisme de Seesaw qui sera utilis par la suite. ### 3.2 Au-del : la violation des nombres leptoniques #### 3.2.1 Les nombres leptoniques Les oscillations entre saveurs de neutrinos remettent en cause la conservation de la saveur leptonique qui est accidentelle dans le MS. Cela n’est pas si choquant car apr s tout, pourquoi n’y aurait-il pas de violations du/des nombre(s) leptonique(s) ? Exp rimentalement, c’est vrai, aucune violation de la saveur leptonique (LFV) n’a t observ e chez les leptons charg s. Par exemple les taux de branchements de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}`$ sont respectivement inf rieurs $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1.2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\times }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{12}$}}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1.1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\times }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{6}$}`$ . Mais th oriquement, il n’y a pas de raison vraiment forte pour justifier la conservation de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}`$, le nombre leptonique total, ou $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}`$, les nombres leptoniques individuels. En physique des particules, les seuls nombres quantiques conserv s sont associ s des sym tries de jauge exactes. De plus, les particules de masses nulles sont aussi associ es de telles sym tries <sup>6</sup><sup>6</sup>6La masse nulle du photon est associ e la sym trie de jauge $`\colorbox[rgb]{1,1,1}{$U$}\colorbox[rgb]{1,1,1}{$($}\colorbox[rgb]{1,1,1}{$1$}\colorbox[rgb]{1,1,1}{$)$}`$ de l’ lectromagn tisme, et celles des gluons la sym trie de jauge de couleur $`\colorbox[rgb]{1,1,1}{$S$}\colorbox[rgb]{1,1,1}{$U$}\colorbox[rgb]{1,1,1}{$($}\colorbox[rgb]{1,1,1}{$3$}\colorbox[rgb]{1,1,1}{$)$}_\colorbox[rgb]{1,1,1}{$C$}`$ de QCD.. Pourtant, il n’y a pas de sym trie de jauge exacte associ e au(x) nombre(s) leptonique(s). Chez les neutrinos, les oscillations entre les diff rentes saveurs montrent bien que les nombres leptoniques individuels ne se conservent pas et que les neutrinos ont des masses non-nulles. Au-del du Mod le Standard en g n ral et dans les mod les supersym triques en particulier, nous nous attendons trouver des LFV dans certains processus <sup>7</sup><sup>7</sup>7La question de la conservation/violation du nombre leptonique total, $`\colorbox[rgb]{1,1,1}{$L$}_\colorbox[rgb]{1,1,1}{$e$}\colorbox[rgb]{1,1,1}{$+$}\colorbox[rgb]{1,1,1}{$L$}_\colorbox[rgb]{1,1,1}{$\mu $}\colorbox[rgb]{1,1,1}{$+$}\colorbox[rgb]{1,1,1}{$L$}_\colorbox[rgb]{1,1,1}{$\tau $}`$, reste cependant ouverte…. Dans cette section, nous discuterons du lien troit entre le secteur des neutrinos et la LFV. #### 3.2.2 Au-del du m canisme de Seesaw ##### Le comptage des param tres Dans le m canisme de Seesaw minimal, le lagrangien du secteur des neutrinos contient : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{`}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{`}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{masse}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{de}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{Dirac}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{"}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{couplage}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{de}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{Yukawa}}$}}{\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\begin{array}{c}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{`}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{`}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{masse}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{de}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{Majorana}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{"}}$}}{\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}{\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}}$$ (3.17) Le m canisme de Seesaw minimal implique 18 param tres physiques. D’un cot les 9 param tres observables basse nergie : les 3 masses l g res, les 3 angles de m lange et les 3 phases violant $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}`$ de la matrice de masse effective basse nergie (4 de ces param tres sont d j connus exp rimentalement, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{12}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{23}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{12}$}}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{23}$}}`$). D’un autre c t , la matrice de masse des neutrinos singulets lourds a aussi 9 degr s de libert , incluant a priori 3 masses lourdes, 3 angles r els et 3 phases CPV. Ceci nous fait un total de 6 phases : la phase de Dirac $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}`$, les 2 phases $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{1,2}}$}}`$ qui affectent la double d sint gration beta sans neutrinos, et 3 phases "lourdes" qui contr lent la leptogen se. Comment pouvons-nous avoir acc s ces 18 param tres ? Les oscillations de neutrinos n’ tant pas suffisantes, quelles sont alors les autres observables que nous pouvons tudier pour d terminer tous les param tres ? ##### La param trisation A haute nergie, le m canisme de Seesaw se param trise par $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ la matrice de couplage de Yukawa des neutrinos droits aux neutrinos gauches (ou par la matrice de masse de Dirac $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{D}$}`$) et par $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ la matrice de masse de Majorana des neutrinos singulets lourds. A basse nergie, nous avons acc s la matrice de masse effective $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}`$ qui est reli e $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ par l’ quation (3.14). Les 9 param tres additionnels n cessaires pour param triser compl tement le m canisme de Seesaw peuvent former une matrice $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\times }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}`$ hermitienne $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$. Nous avons alors sch matiquement : $$\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}}{\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}}{\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (3.18) Reste savoir comment obtenir des informations sur $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ partir de mesures d’observables basses nergies et comment relier $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Il se trouve que dans le mod le supersym trique seesaw minimal, les processus violant le nombre leptonique permettent justement l’ tude indirecte des param tres du secteur des neutrinos car ils donnent acc s cette matrice $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$. #### 3.2.3 Les processus violant la saveur leptonique Dans un sc nario seesaw minimal et supersym trique, certains processus qui violent la saveur leptonique peuvent tre induits par l’interaction de Yukawa $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$. En effet, si la supersym trie se brise des nergies plus grandes que l’ chelle de masse des neutrinos lourds, des termes non-diagonaux des matrices de masse des sleptons permettant la violation de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}`$ et de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}`$ seront induites radiativement gr ce aux couplages de Yukawa des neutrinos l gers, et ceci m me si les param tres d’origine de la brisure sont ind pendants de la saveur <sup>8</sup><sup>8</sup>8Si on suppose que les termes softs sont r els et universels l’ chelle de l’unification, comme sugg r par l’absence de grandes contributions supersym triques aux processus LFV et CPV, les violations de $`\colorbox[rgb]{1,1,1}{$C$}\colorbox[rgb]{1,1,1}{$P$}`$ et des nombres leptoniques dans le secteur des sleptons sont enti remement induites par les effets de renormalisation des param tres de couplage des neutrinos.. C’est la renormalisation de la masse "douce" des sleptons qui va cr er le m lange des saveurs. A l’ordre d’une boucle et quand $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}`$, la renormalisation des param tres de brisure de supersym trie basse nergie se trouve tre proportionnelle : $$\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{ln}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (3.19) o $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}}`$ est l’ chelle o on impose les conditions initiales sur les param tres de brisure de supersym trie (universalit des masses scalaires $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}}`$ …). Les $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}`$ sont les masses de neutrinos lourds. Or (3.19) peut exactement jouer le r le de la matrice $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ . En particulier, nous avons les renormalisations suivantes : $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{8}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{8}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}`$ (3.20) Dans ce cas, la seule combinaison des angles et des phases qui viole $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}`$, analogue l’invariant de Jarlskog dans le MS <sup>9</sup><sup>9</sup>9Il traduit l’amplitude de la violation de $`\colorbox[rgb]{1,1,1}{$C$}\colorbox[rgb]{1,1,1}{$P$}`$ dans le secteur des quarks. est: $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{Im}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{[}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{12}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{23}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{31}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{]}$}$$ (3.21) et elle ne d pend que d’une seule phase. Remarque : si, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}}`$, les valeurs initiales des param tres de brisure douce ne sont pas ind pendantes de la saveur, il y aura des sources additionnelles aux processus changeant la saveur en dehors de celles discut es ici. Seconde remarque, si les neutrinos lourds ne sont pas d g n r s, il y a d’autres contributions $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}}}`$ qui d pendront alors de 3 phases . Tout ceci peut conduire des taux de branchements observables pour les processus LFV comme $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}`$, la conversion $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}`$ dans les noyaux, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}`$. En g n ral, comme $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ est complexe, elle conduit une violation de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}`$ dans les oscillations de neutrinos, dans les processus rares LFV ainsi que dans les moments dipolaires lectriques. Cette violation de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}`$ leptonique est primordiale pour que l’asym trie baryonique observ e dans l’Univers ait une origine dans la leptogen se. Nous allons voir des exemples num riques de tous ces cas. ##### La LFV chez les leptons charg s Nous supposons dans les exemples num riques suivants que les seules sources de LFV et CPV sont d es aux interactions avec les neutrinos singulets lourds. Si nous calculons cette d sint gration dans le mod le seesaw supersym trique et minimal, (avec un choix particulier des param tres des neutrinos) et pour une grande partie de l’espace des param tres supersym triques, la pr diction est au-dessus de la limite exp rimentale. Mais quand la pr diction du taux de branchement est minimale (i.e. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{300}$}`$ GeV, figure (3.6); $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2.1}$}`$, figure (3.7)) l’asym trie $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒜}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}`$ de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}`$ qui viole $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}`$ (donc de fa on quivalente $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}`$) est maximale (elle peut tre de 10$`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\%}$}`$ environ) : il y a anti-corr lation entre $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒜}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Le taux de branchement $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}`$ est donc dans ce cas comparable celui de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}`$. Ceci est en fait d des annulations possibles dans les diff rents diagrammes pingouins photoniques. Ce r sultat est commun d’autres choix sur les param tres des neutrinos, nous n’entrerons donc pas dans les d tails. Ainsi, il est en principe possible de mesurer exp rimentalement la violation de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}`$ dans le secteur des leptons charg s avec la d sint gration $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}`$. La violation de la saveur leptonique peut aussi se voir dans les d sint grations de taus, par exemple $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}`$ ou $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}`$. Mais pour $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}`$, il n’y a pas de telles annulations dans la r gion de param tres consid r e dans les figures (3.8). Dans ce sc nario, la plupart des points se situent un peu en dessous des limites exp rimentales actuelles et donc les taux de branchements pourraient, en principe, avoir des chances d’ tre observables au LHC ou Babar et Belle. ##### Les moments dipolaires lectriques leptoniques Les violations de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}`$ et des saveurs leptoniques ont aussi un impact sur les moments dipolaires lectriques (EDM) des leptons. Ceux-ci d pendent en principe de phases qui apparaissent aussi dans la leptogen se. Les valeurs num riques de ces EDM sont augment s de plusieurs ordres de grandeur quand les neutrinos lourds sont non-d g n r s. Or dans la plupart des mod les de masse des neutrinos ph nologiquement viables, les neutrinos lourds ne sont pas d g n r s. L’EDM d’un lepton $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}`$ est d fini comme tant le coefficient $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}`$ de l’interaction suivante : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (3.22) Dans le MSSM, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}`$ re oit des contributions des boucles de neutralinos et de charginos et d pend fortement de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$. Les contributions dominantes $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}`$ sont dues aux termes suivants <sup>10</sup><sup>10</sup>10Ce sont en fait les m me termes que l’ quation (3.20) mais dans le cas de neutrinos lourds non d g n r s. de la renormalisation des termes de brisure douce: $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ $`{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{18}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{ln}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{11}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{7}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{[}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{]}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{ln}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}`$ (3.23) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\frac{\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}`$ et $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\sqrt{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}}}`$. $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}`$ est la masse moyenne des neutrinos singulets lourds (qui ne sont pas d g n r s dans ce cas). Les pr dictions des EDM leptoniques sont illustr es num riquement dans la figure (3.9). Les EDM du muon et de l’ lectron peuvent atteindre $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{27}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{28}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}`$.cm et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{29}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{30}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}`$.cm respectivement. Les limites sup rieures actuelles sont : $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4.3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\times }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{27}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{e}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cm}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3.7}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3.4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\times }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{19}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{e}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cm}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3.1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\times }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{16}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{e}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{cm}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}`$ (3.24) Les exp riences futures BNL et aupr s d’usines neutrinos projettent respectivement de sonder l’EDM du muon jusqu’ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{24}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}`$.cm et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{26}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}`$.cm. La mesure de l’EDM de l’ lectron devrait atteindre les environs de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{33}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}`$.cm. Ces exp riences ont donc la possibilit de tester le mod le seesaw supersym trique minimal via ses pr dictions sur $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}`$ ($`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}`$ risque malheureusement d’ tre difficile mesurer). #### 3.2.4 La d sint gration de sparticules Une autre fa on de mesurer la violation du/des nombres leptoniques est d’observer la d sint gration de sparticules. Par exemple, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ o $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ est la LSP et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ la particule la plus l g re apr s $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ (NLSP). En particulier, les d sint grations de sparticules qui violent $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}`$ sont les plus int ressantes car les effets de la renormalisation des param tres de brisure douce sont plus importants dans le cas du $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}`$ (le couplage de Yukawa du $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}`$ est plus grand). Cela rend les effets de violation du nombre leptonique $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}`$ potentiellement grands pour les d sint grations comme $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}`$ ou $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}`$. Les recherches au LHC de ces d sint grations pourraient donc tre compl mentaires celles de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}`$. Dans la figure suivante, (3.10), les lignes en traits pleins d notent les largeurs de d sint gration de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ en diff rents modes, dans le cas particulier du CMSSM <sup>11</sup><sup>11</sup>11En fait, une non-universalit des masses scalaires a t introduite, $`\colorbox[rgb]{1,1,1}{$($}\colorbox[rgb]{1,1,1}{$m$}_\colorbox[rgb]{1,1,1}{$0$}^\colorbox[rgb]{1,1,1}{$2$}\colorbox[rgb]{1,1,1}{$)$}_{\colorbox[rgb]{1,1,1}{$L$}\colorbox[rgb]{1,1,1}{$L$}}\colorbox[rgb]{1,1,1}{$=$}\colorbox[rgb]{1,1,1}{$d$}\colorbox[rgb]{1,1,1}{$i$}\colorbox[rgb]{1,1,1}{$a$}\colorbox[rgb]{1,1,1}{$g$}\colorbox[rgb]{1,1,1}{$($}\colorbox[rgb]{1,1,1}{$m$}_\colorbox[rgb]{1,1,1}{$0$}^\colorbox[rgb]{1,1,1}{$2$}\colorbox[rgb]{1,1,1}{$,$}\colorbox[rgb]{1,1,1}{$m$}_\colorbox[rgb]{1,1,1}{$0$}^\colorbox[rgb]{1,1,1}{$2$}\colorbox[rgb]{1,1,1}{$,$}\colorbox[rgb]{1,1,1}{$x$}\colorbox[rgb]{1,1,1}{$\times $}\colorbox[rgb]{1,1,1}{$m$}_\colorbox[rgb]{1,1,1}{$0$}^\colorbox[rgb]{1,1,1}{$2$}\colorbox[rgb]{1,1,1}{$)$}`$ avec $`\colorbox[rgb]{1,1,1}{$x$}\colorbox[rgb]{1,1,1}{$$}\colorbox[rgb]{1,1,1}{$0.9$}`$. De plus, l’angle de m lange dans la matrice de Yukawa entre la 2 me et 3 me g n ration de lepton charg s, $`\colorbox[rgb]{1,1,1}{$\varphi $}`$, a t pris gal $`\colorbox[rgb]{1,1,1}{$\pi $}\colorbox[rgb]{1,1,1}{$/$}\colorbox[rgb]{1,1,1}{$6$}`$ dans les 2 exemples. avec les param tres $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{600}$}`$ GeV. Le rapport des taux de branchements $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ est repr sent par les tirets. Nous voyons que pour $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{270}$}`$ GeV, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ est de l’ordre de 1 donc $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ peut tre comparable $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. La violation du nombre leptonique peut donc tre importante. De plus, sur la figure (3.11) les contours des taux de branchements sont trac s pour le m me choix de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{tan}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}`$ mais dans le plan $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. ### 3.3 Les neutrinos et la cosmologie #### 3.3.1 L’asym trie baryonique La physique des neutrinos peut aussi jouer un r le en cosmologie, particuli rement pour r pondre la question de l’origine de la mati re. En effet, la quasi-totalit de ce que nous observons dans l’Univers est fait de mati re, nous sommes fait de mati re, mais pas d’antimati re. Pourquoi cette asym trie ? La densit baryonique de l’Univers est : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\eta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1.5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{6.3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\times }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (3.25) o $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}`$ sont respectivement le nombre de baryons et le nombre de photons. L’asym trie baryonique cosmologique est alors $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}_{\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}`$, o $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}`$ est la densit d’entropie, et vaut environ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\eta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{7}$}`$. Cette asym trie peut s’expliquer <sup>12</sup><sup>12</sup>12La possibilit que nous ne vivions que dans une r gion dans l’Univers constitu e de mati re et que l’antimati re serait localis hors de notre horizon est tr s d favoris e. par un tr s faible d s quilibre dans l’Univers primordial en faveur de la mati re, caus par les interactions. Par la suite, l’antimati re s’est alors annihil e avec la mati re et le reste a form un peu plus tard les baryons dont nous sommes faits. Mais une asym trie mati re-antimati re ne peut tre dynamiquement g n r e dans un univers en expansion que si les interactions entre particules et l’ volution cosmologique satisfont les 3 conditions de Sakharov : * violation du nombre baryonique ($`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ les interactions changent alors le nombre de quarks) * violation de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}`$ ($`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ les interactions sont donc diff rentes pour la mati re et l’antimati re) * d viation de l’ quilibre thermique, par exemple lors d’une transition de phase ($`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ vite que les produits de d sint gration ne puisse recr er la particule initiale et donc permet de cumuler le d s quilibre) La question de l’asym trie baryonique permet donc d’ tablir un lien entre le mod le standard cosmologique et le mod le standard de la physique des particules. A noter qu’une asym trie g n r e une poque de l’Univers peut tre effac e par la suite si elle n’est pas prot g e, ce qu’il faut viter. Cependant, nous ne discuterons pas de ce genre de d tails (bien que ce soit important). #### 3.3.2 La leptogen se Le m canisme de g n ration de cette asym trie baryonique est appel e la baryogen se. Il existe dej plusieurs sc nario viables. Mais ce dont nous allons discuter ici est la leptogen se, dans laquelle l’asym trie baryonique est obtenue via une asym trie leptonique dans la d sint gration de neutrinos singulets lourds (les neutrinos droits de Majorana), ce qui est possible si le nombre leptonique n’est pas conserv . La d sint gration $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}`$ se fait avec un taux de branchement diff rent de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}}`$. L’asym trie entre le nombre de leptons et le nombre d’antileptons est convertie en une asym trie entre le nombre de quarks et d’antiquarks par des interactions lectrofaibles non-perturbatives (les sphal rons) qui donneraient : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (3.26) avec $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{8}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{23}$}`$ dans le MSSM. Le taux de d sint gration total d’un neutrino singulet lourd $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ peut s’ crire ainsi (sans sommation sur i) : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{8}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pi }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (3.27) Les diagrammes une boucle qui impliquent l’ change d’un neutrino lourd, figure (3.12), peuvent g n rer une asym trie $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}`$ dans la d sint gration des neutrinos lourds $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$. Cette asym trie s’ crit sous la forme suivante : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{8}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pi }$}}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{Im}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{[}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{]}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\times }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (3.28) o $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}`$ est une fonction cin matique connue et calculable et $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (3.29) Nous voyons dans (3.28) que la leptogen se est proportionnelle au produit $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}`$ qui d pend de 9 param tres r els et de 3 phases CPV mais pas celles de basse nergie ($`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{1,2}}$}}`$). L’existence d’une asym trie leptonique ne requiert donc pas que $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}`$ soit non-nulle. Sur la figure (3.13), nous voyons que les cas $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ sont indistinguables. Comment avoir acc s la leptogen se ? Il est possible de formuler une "strat gie" pour calculer la leptogen se en terme d’observables mesurables en laboratoire : * mesurer la phase $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}`$ des oscillations de neutrinos et les phases de Majorana $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{1,2}}$}}`$, * mesurer les observables reli es la renormalisation des param tres de brisure douce de la supersym trie qui sont fonctions de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{1,2}}$}}`$ et des phases de la leptogen se, * extraire les effets connus de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{1,2}}$}}`$ pour isoler les param tres de la leptogen se. A l’heure actuelle il nous manque les informations sur les 2 premi res tapes. Nous pouvons juste explorer l’espace des param tres. Pour chaque ($`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}`$) et ($`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}`$), il est possible de calculer l’observable $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}}`$ de la double d sint gration beta sans neutrinos, les EDM, les processus LFV et l’asym trie $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$. Il est aussi possible d’observer les corr lations entre tous ces param tres et les d pendances aux masses l g res et lourdes des neutrinos. Les figures suivantes, (3.14) et (3.15), montrent qu’il est possible d’obtenir le bon nombre de baryons partir de la physique des neutrinos et de la physique de basse nergie, en observant les moments dipolaires lectriques des leptons charg s, les d sint grations rares $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}`$, etc. De plus, il est possible d’obtenir des renseignements sur l’espace des param tres de haute nergie comme les masses $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}`$. La limite $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\times }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{11}$}}`$ implique une limite sur la masse du neutrino lourd le plus l ger : $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{GeV}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{o}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{11}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{GeV}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{h}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ (3.30) Mais ceci peut potentiellement entrer en conflit avec la limite inf rieure non-triviale d e une possible surproduction de gravitinos. Cette surproduction modifierait alors les abondances des l ments produits au cours de la nucl osynth se primordiale (He, Li,…). La leptogen se est un sc nario important prendre en compte et montre tout l’int r t de consid rer les liens entre la physique des neutrinos et la cosmologie. #### 3.3.3 L’inflation sneutrinique Un peu plus en marge du cours, mais n anmoins int ressante, est la possibilit que l’inflaton, champ scalaire responsable de l’inflation de l’Univers, soit incarn par l’un des partenaires supersym triques des neutrinos singulets lourds (voir par exemple la r f rence ). ##### Un br ve interlude sur l’inflation Nous allons tout d’abord commencer par quelques mots sur la cosmologie et l’inflation. A l’origine de l’hypoth se de l’inflation il y a les probl mes auxquels sont confront s les cosmologistes : pourquoi l’Univers semble si homog ne ? Pourquoi est-il si vieux ou si vaste ? Pourquoi est-il si plat ? Pourquoi l’entropie de l’Univers est-elle si lev e ? Pour rem dier cel , nous pouvons introduire l’id e d’une p riode inflationnaire c’est- -dire l’id e qu’ une poque de son histoire, l’expansion de l’Univers fut quasi-exponentielle. Cette expansion exponentielle agrandit alors consid rablement les dimensions de l’Univers et r duit la densit d’ nergie (et par cons quent la temp rature globale). Ceci est rendu possible par l’introduction d’un champ scalaire, l’inflaton, qui, au tout d but de l’Univers, contenait la quasi-totalit de l’ nergie. La figure (3.16) montre une forme possible du potentiel scalaire de l’inflaton et les diff rentes tapes de l’inflation. A la premi re tape, le potentiel de l’inflation part d’une valeur initiale et "roule" jusqu’au puit : l’Univers subit une expansion inflationnaire exponentielle. Ensuite, quand l’inflaton arrive dans le puit du potentiel, il y passe un certain temps en oscillant jusqu’ se stabiliser. Ceci correspond au moment o l’inflaton se d sint gre en mati re par la conversion de l’ nergie initialement sous forme d’inflaton en photons et autres particules l g res. Ainsi l’univers est r chauff jusqu’à une temp rature de r chauffement $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}}`$. L’inflation donne aussi une origine aux petites fluctuations de densit observ es par WMAP (mais d j vues par COBE), qui sont ind pendantes de l’ chelle de la structure : elles viendraient de fluctuations quantiques de la valeur du potentiel pendant l’inflation. Il existe plusieurs mod les d’inflation, mais l’id e pr sent e reste la m me. Cependant, personne ne sait ce qu’est l’inflaton et quelle particule il peut tre identifi . ##### Le sneutrino, un candidat potentiel Dans le cas particulier d’un mod le d’inflation chaotique <sup>13</sup><sup>13</sup>13C’est un sc nario inflationnaire dans lequel il n’y a pas de structure privil gi e dans le potentiel $`\colorbox[rgb]{1,1,1}{$V$}\colorbox[rgb]{1,1,1}{$($}\colorbox[rgb]{1,1,1}{$\varphi $}\colorbox[rgb]{1,1,1}{$)$}`$. Il peut tre une simple puissance, $`\colorbox[rgb]{1,1,1}{$V$}\colorbox[rgb]{1,1,1}{$$}\colorbox[rgb]{1,1,1}{$\varphi $}^\colorbox[rgb]{1,1,1}{$n$}`$ ou exponentiel, $`\colorbox[rgb]{1,1,1}{$V$}\colorbox[rgb]{1,1,1}{$$}\colorbox[rgb]{1,1,1}{$e$}^{\colorbox[rgb]{1,1,1}{$\alpha $}\colorbox[rgb]{1,1,1}{$\varphi $}}`$. Une r gion donn e de l’Univers est suppos e commencer avec une valeur particuli re de $`\colorbox[rgb]{1,1,1}{$\varphi $}`$, donc de $`\colorbox[rgb]{1,1,1}{$V$}`$, qui d croit ensuite de fa on monotone et lente vers z ro (c’est le ”slow-roll”). avec un potentiel de la forme $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$, les donn es de WMAP donnent pour la masse de l’inflaton : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{I}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\times }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{13}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{GeV}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (3.31) Or cette masse correspond plus ou moins l’intervalle de masse consid r pour les neutrinos singulets lourds (et a priori aussi pour les sneutrinos associ s) : $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{15}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{GeV}}$}`$. De plus, l’inflaton est sens ne pas avoir d’interaction de jauge, ce qui peut tr s bien tre le cas des sneutrinos lourds. Si un de ces sneutrinos est l’inflaton, alors le probl me cosmologique de l’inflation sera connect au reste de la physique des particules. De plus, cette hypoth se peut tre utilis e pour donner des pr dictions sur les d sint grations violant le nombre leptonique car elle contraint significativement les param tres du mod le seesaw supersym trique minimal. Dans la figure (3.17), nous voyons entre autres que le processus $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}`$ est tr s proche de la limite actuelle et dans les possibilit s observationnelles des futures exp riences. Nous observons aussi que les taux de branchement de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}`$ semblent ind pendants de la temp rature de r chauffement (pour $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{12}$}}`$ GeV) et que $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ est tr s sensible $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{sin}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{13}$}}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}}`$. ### 3.4 Conclusions Le m lange des neutrinos observ via les oscillations peut avoir de nombreuses cons quences, notamment dans un cadre supersym trique o la masse si faible des neutrinos tire son origine du m canisme de Seesaw. Par exemple, les taux de branchement des processus violant la saveur leptonique et les moments dipolaires lectriques leptoniques calcul s dans ce mod le seesaw supersym trique minimal montrent que la violation des nombres leptoniques et la violation de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}`$ leptonique peuvent avoir lieu de mani re significative. Les futures exp riences devraient avoir la possibilit de mesurer pour la premi re fois ces violations. Ceci donne alors de grandes chances de pouvoir reconstruire les param tres de haute nergie de la physique des neutrinos dans ce mod le, inaccessibles depuis les exp riences qui observent les oscillations des neutrinos (figure (3.18)). De plus, la physique des neutrinos permet aussi de r soudre certains probl mes de nature cosmologique gr ce aux hypoth ses telles que la leptogen se et l’inflation sleptonique. ## Chapitre 4 Le grand "Au-del " ### 4.1 La Grande Unification Les th ories de jauge (surtout non-ab liennes, appel es th ories de Yang-Mills) semblent offrir un cadre unique pour d crire la physique des particules. Dans le MS, il y a 3 couplages diff rents (si la gravitation n’est pas compt e) associ s aux trois groupes de jauge $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}`$ et certains param tres sont ajust s arbitrairement pour rendre compte des diverses observations exp rimentales. Un prolongement logique du MS est de consid rer l’existence plus haute nergie d’une sym trie plus grande. Celle-ci pourrait permettre de relier les diff rents param tres et d’unifier les 3 couplages <sup>1</sup><sup>1</sup>1On note les couplages ainsi : $`\colorbox[rgb]{1,1,1}{$g$}_\colorbox[rgb]{1,1,1}{$1$}`$ pour le couplage de $`\colorbox[rgb]{1,1,1}{$U$}\colorbox[rgb]{1,1,1}{$($}\colorbox[rgb]{1,1,1}{$1$}\colorbox[rgb]{1,1,1}{$)$}`$, $`\colorbox[rgb]{1,1,1}{$g$}_\colorbox[rgb]{1,1,1}{$2$}`$ pour $`\colorbox[rgb]{1,1,1}{$S$}\colorbox[rgb]{1,1,1}{$U$}\colorbox[rgb]{1,1,1}{$($}\colorbox[rgb]{1,1,1}{$2$}\colorbox[rgb]{1,1,1}{$)$}`$ et $`\colorbox[rgb]{1,1,1}{$g$}_\colorbox[rgb]{1,1,1}{$3$}`$ pour $`\colorbox[rgb]{1,1,1}{$S$}\colorbox[rgb]{1,1,1}{$U$}\colorbox[rgb]{1,1,1}{$($}\colorbox[rgb]{1,1,1}{$3$}\colorbox[rgb]{1,1,1}{$)$}`$. Dans les chapitres pr c dents on avait $`\colorbox[rgb]{1,1,1}{$g$}`$ pour $`\colorbox[rgb]{1,1,1}{$g$}_\colorbox[rgb]{1,1,1}{$2$}`$ et $`\colorbox[rgb]{1,1,1}{$g$}^{\colorbox[rgb]{1,1,1}{$$}}`$ pour $`\colorbox[rgb]{1,1,1}{$g$}_\colorbox[rgb]{1,1,1}{$Y$}`$. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}`$. La philosophie de la Grande Unification (GU) est justement de chercher un groupe de jauge G simple qui comprend $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}`$. Ainsi, les 3 forces fondamentales ( lectromagn tique, faible, forte) se trouveraient unifi es en une force " lectronucl aire" de couplage unique $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}}`$. Ceci semble du point de vue le plus courant en physique des particules plus "satisfaisant" car cela va dans le sens d’une physique plus "simple" aux chelles de longueurs plus petites. Cependant, les effets de l’interaction gravitationnelle sont suppos s tre toujours n gligeables, au moins en premi re approximation. Si l’ chelle d’unification $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}}`$ se trouve tre significativement plus petite que la masse de Planck alors cette supposition est justifi e. Il se trouve que les estimations typiques, bas es sur une extrapolation tr s haute nergie de la physique connue (c’est- -dire le MS), donnent une chelle de l’ordre de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{16}$}}`$ GeV, environ mille fois plus petite que l’ chelle de Planck $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒪}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{19}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ GeV. Le fait d’avoir un unique groupe de jauge pour d crire la physique cette chelle implique aussi des relations tout fait in dites entre les particules, ainsi que de nouveaux bosons de jauge. En effet, la sym trie change donc l’organisation des particules (les multiplets) change. Certains indices comme la quantification de la charge (i.e. l’existence de charges lectriques fractionnaires) ou la compensation des anomalies <sup>2</sup><sup>2</sup>2C’est- -dire le fait que les anomalies dues aux leptons s’annulent exactement avec celles dues aux quarks. font aussi penser une organisation plus simple que celle du MS. Bien s r, basse nergie, nous devons retrouver le Mod le Standard, et dans ces Th ories de Grande Unification (GUT) nous devrons donc tudier la brisure du groupe de jauge G en $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}`$. Cette section commence donc par une pr sentation des quations d’ volution des 3 couplages et de leur unification. Nous exposerons ensuite quelques exemples de mod les dont le "prototype" bas sur le groupe $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ permet d’aborder beaucoup de propri t s des GUT sans alourdir la discussion. Puis nous tudierons les pr dictions typiques de ces mod les comme la d sint gration du proton et les relations entre les masses des quarks et leptons. Nous terminerons en discutant des avantages, des probl mes et des perspectives des mod les bas s sur l’id e d’une Grande Unification haute nergie. #### 4.1.1 Les quations d’ volution des couplages $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ ##### L’unification des couplages L’un des r sultats les plus importants de la renormalisation est l’ volution des constantes de couplage des forces en fonction de l’ chelle d’ nergie : $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. En supposant que le MS reste valable jusqu’ des nergies de l’ordre de la masse de Planck nous pouvons r soudre analytiquement ces quations du groupe de renormalisation (RGE) et observer l’ volution des 3 couplages des 3 interactions fondamentales du MS. Le r sultat, surprenant, initialement obtenu dans les ann es 70 est que les couplages voluent logarithmiquement (nous justifierons un peu plus tard cette remarque) jusqu’ se croiser au m me point (à peu près) et donc atteindre une valeur commune. Cette observation l gitimise la philosophie de la Grande Unification. En r alit , depuis la premi re fois o le calcul a t fait, la pr cison exp rimentale a grandement augment et les couplages ne se croisent plus tout- -fait au m me point. La figure (4.1) montre clairement ce "rendez-vous" manqu . Cela veut-il dire que nous devons oublier la GU ? Non, simplement parce que nous ne connaissons pas toute la physique au-del du MS et que pour obtenir la figure (4.1) nous avons fait une supposition qui n’ tait sans doute pas justifi e : le MS est valable jusqu’ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}}`$. Au chapitre 1 nous nous sommes justement efforc s de montrer que le MS ne pouvait pas tre valable au-del d’une chelle d’ nergie, de l’ordre du TeV. Il y a de la nouvelle physique aux alentours de cette chelle et celle-ci peut changer drastiquement les quations d’ volution des constantes de couplage. De plus, au chapitre 2 nous avons vu que la supersym trie tait une candidate tr s s rieuse pour cette nouvelle physique et celle-ci introduit un grand nombre de nouvelles particules qui interviennent dans la renormalisation des couplages. Le calcul des RGE dans le cadre du MSSM donne alors la figure (4.2). Le r sultat est tr s surprenant lui aussi puisque nous retrouvons l’unification des couplages ! La supersym trie semble donc un cadre appropri pour la GU. Pour "mesurer la qualit " de cette unification, on adopte souvent un point de vue inverse en partant de l’unification de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$, en supposant l’unification avec $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}`$ exacte et au m me point, puis en calculant alors l’ volution de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}`$ vers les basses nergies <sup>3</sup><sup>3</sup>3En effet, les 3 couplages de basse nergie sont maintenant fonction de 2 param tres ind pendants, $`\colorbox[rgb]{1,1,1}{$\alpha $}_{\colorbox[rgb]{1,1,1}{$G$}\colorbox[rgb]{1,1,1}{$U$}}`$ et $`\colorbox[rgb]{1,1,1}{$M$}_{\colorbox[rgb]{1,1,1}{$G$}\colorbox[rgb]{1,1,1}{$U$}}`$, il y a donc une pr diction possible.. Nous pouvons alors comparer la valeur de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}`$ ( $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$ par exemple) avec la valeur exp rimentale et v rifier la compatibilit . Dans le MSSM, pour une unification $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\times }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{16}$}}`$ GeV, nous trouvons alors la valeur : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0.130}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0.004}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (4.1) o $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}}`$ sont des corrections qui d pendent du mod le. Les donn es exp rimentales donnent la valeur suivante : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0.117}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pm }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0.002}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (4.2) qui est en assez bon accord avec la pr diction. Sans les corrections $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}}`$, l’accord ne serait pas parfait, ce qui voudrait dire que l’unification haute nergie des 3 couplages ne serait pas tout fait exacte. Mais il ne faut pas oublier que dans le calcul des RGE nous nous sommes arr t s une certaine pr cision ( un ordre donn e de la th orie des perturbations). La correction $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}}`$ est un effet de seuil basse nergie, venant de l’ chelle de brisure effective (douce) de la supersym trie. En plus, si effectivement il y a une Grande Unification et donc des nouvelles particules aux hautes énergies, celles-ci devaient se faire "sentir" quand nous nous rapprochons de l’ chelle $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}}`$. Plus pr cisemment, aux chelles d’ nergie proches de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}}`$, les boucles virtuelles de ces bosons ont un effet non n gligeable, car les masses de ces bosons ne sont plus n gligeables devant l’ chelle d’ nergie o l’on se trouve. On parle d’effet de seuil GUT, repr sent par la quantit $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}}`$ de l’ quation 4.1. Malheureusement, il d pend de la physique l’ chelle $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}}`$, et pour calculer quantitativement sa valeur il faut se placer dans un mod le donn . On parle de "sensibilit la th orie UV" <sup>4</sup><sup>4</sup>4UV pour ultra-violette, c’est- -dire de plus haute nergie. de l’unification des couplages. Une remarque importante : par rapport au MS, dans un mod le GUT le couplage de l’hypercharge a subit une modification d’un facteur $`\sqrt{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}}`$ ; c’est en r alit la relation $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\sqrt{\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}}`$ que nous avons l’ chelle $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}}`$. Nous verrons un peu plus tard pourquoi ce facteur a t introduit. Maintenant, nous allons tre un peu plus technique et donner ces RGE dans les cas discut s savoir le MS et le MSSM. ##### Les RGE Les quations du groupe de renormalisation pour les trois couplages des trois interactions non-gravitationnelles se pr sentent une boucle sous la forme : $$\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{d}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{d}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{16}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{1,2,3}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (4.3) o $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{ln}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ avec $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}`$ une chelle d’ nergie. Nous utiliserons plus souvent $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ la place des $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ o $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}}`$. Les RGE s’ crivent alors : $$\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{d}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{d}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pi }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{1,2,3}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (4.4) Le tableau 4.1 donne les diff rentes valeurs de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ dans le cas du MS (un seul doublet de Higgs et 3 g n rations) et du MSSM. Les expressions o le nombre de doublets de Higgs et de familles de quarks sont gard s explicites peuvent se trouver dans . Ces quations diff rentielles peuvent se r soudre analytiquement et les solutions s’ crivent ainsi : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pi }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{ln}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}$$ (4.5) Avec ces solutions, nous pouvons donc maintenant v rifier quelle pr cision et quelle chelle les couplages s’unifient, dans le MS et le MSSM. En g n ral, on utilise la pr diction de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{sin}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ ou celle de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$. Nous n’introduisons aucunes corrections dûes aux effets de seuil d’un mod le GUT particulier ou aux effets de seuil basse nergie et nous supposons la relation $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}}`$. En prenant comme chelle $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}`$ l’ chelle $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$, $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{sin}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (4.6) et en utilisant les relations <sup>5</sup><sup>5</sup>5Le facteur 5/3 vient de la normalisation du g n rateur de la charge lectrique dans les mod les GU. Dans le mod le standard, $`\colorbox[rgb]{1,1,1}{$Q$}\colorbox[rgb]{1,1,1}{$=$}\colorbox[rgb]{1,1,1}{$T$}_\colorbox[rgb]{1,1,1}{$3$}\colorbox[rgb]{1,1,1}{$+$}\colorbox[rgb]{1,1,1}{$Y$}\colorbox[rgb]{1,1,1}{$/$}\colorbox[rgb]{1,1,1}{$2$}`$. Dans les GUTs, $`\colorbox[rgb]{1,1,1}{$Q$}\colorbox[rgb]{1,1,1}{$=$}\frac{\colorbox[rgb]{1,1,1}{$1$}}{\colorbox[rgb]{1,1,1}{$2$}}\colorbox[rgb]{1,1,1}{$($}\colorbox[rgb]{1,1,1}{$T$}^{\colorbox[rgb]{1,1,1}{$11$}}\colorbox[rgb]{1,1,1}{$+$}\sqrt{\frac{\colorbox[rgb]{1,1,1}{$5$}}{\colorbox[rgb]{1,1,1}{$3$}}}\colorbox[rgb]{1,1,1}{$T$}^{\colorbox[rgb]{1,1,1}{$12$}}\colorbox[rgb]{1,1,1}{$)$}`$ o $`\colorbox[rgb]{1,1,1}{$T$}^{\colorbox[rgb]{1,1,1}{$11$}}`$ et $`\colorbox[rgb]{1,1,1}{$T$}^{\colorbox[rgb]{1,1,1}{$12$}}`$ sont les deux g n rateurs de $`\colorbox[rgb]{1,1,1}{$S$}\colorbox[rgb]{1,1,1}{$U$}\colorbox[rgb]{1,1,1}{$($}\colorbox[rgb]{1,1,1}{$2$}\colorbox[rgb]{1,1,1}{$)$}_\colorbox[rgb]{1,1,1}{$L$}\colorbox[rgb]{1,1,1}{$$}\colorbox[rgb]{1,1,1}{$U$}\colorbox[rgb]{1,1,1}{$($}\colorbox[rgb]{1,1,1}{$1$}\colorbox[rgb]{1,1,1}{$)$}_\colorbox[rgb]{1,1,1}{$Y$}`$ du groupe d’unification. L’isopin et l’hypercharge font partie d’un seul groupe simple $`\colorbox[rgb]{1,1,1}{$G$}`$ et ne sont pas ind pendants. L’hypercharge est alors red finie par un facteur $`\sqrt{\frac{\colorbox[rgb]{1,1,1}{$5$}}{\colorbox[rgb]{1,1,1}{$3$}}}`$ que l’on r percute sur le couplage $`\colorbox[rgb]{1,1,1}{$g$}_\colorbox[rgb]{1,1,1}{$Y$}`$. $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pi }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pi }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (4.7) et la combinaison particuli re : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pi }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{ln}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (4.8) on trouve dans le MS: $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{sin}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{8}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{55}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{24}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pi }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{ln}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (4.9) et dans le MSSM : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{sin}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{8}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{7}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pi }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{ln}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (4.10) A noter que dans le cas du MSSM, nous ne connaissons pas les masses des sparticules. En les mettant toutes 1 TeV on retrouve approximativement la valeur de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{sin}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ observ e exp rimentalement. Pour examiner ces pr dictions avec plus de d tails nous pouvons tudier les quations du groupe de renormalisation jusqu’ l’ordre de 2 boucles. Dans le MSSM, l’unification est cependant stable car les corrections 2 boucles sont faibles devant celles 1 boucle, m me si cela reste moins vrai pour $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}`$. Le r sultat est une unification encore plus pr cise c’est- -dire une pr diction de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{sin}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ plus proche de la valeur exp rimentale. On peut retrouver les expressions 2 boucles et plus de d tails sur les effets de seuil des masses des sparticules dans . #### 4.1.2 Les mod les Grand Unifi s ##### Le choix du groupe d’unification Quels sont les groupes susceptibles de nous int resser pour construire une th orie Grande Unifi e? Tout d’abord, ces groupes sont des groupes dits "de Lie" suffisamment "grands" pour englober enti rement celui du Mod le Standard. Ce dernier est de rang 4, c’est- -dire que le nombre de g n rateurs de la sym trie simultan ment diagonalisables <sup>6</sup><sup>6</sup>6Chacun est associ un nombre quantique, une ”charge”, et les tats de particules observ s sont index s par ceux-ci. A noter que ces g n rateurs diagonaux sont de trace nulle. est gal 4 : $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}`$ en contient 2, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}`$ un seul et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}`$ aussi. Il nous faut donc trouver dans la classification des groupes de Lie, faite par Cartan, un groupe de rang sup rieur ou gal 4. Enfin, ils doivent comporter des repr sentations complexes pour que les fermions de chiralit diff rentes puissent tre dans des repr sentations diff rentes. Les deux possibilit s de rang 4 sont: $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Mais $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ ne permet pas simultan ment aux leptons d’avoir une charge lectrique enti re et aux quarks d’avoir une charge lectrique fractionnaire. Le groupe $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ est donc le groupe le plus simple possible capable d’englober le Mod le Standard. Les autres groupes possibles et couramment utilis s mais de rangs sup rieurs 4 sont $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{O}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, de rang 5, et le groupe exceptionnel $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{6}$}`$, de rang 6. Nous le voyons, les consid rations physiques pr c dentes contraignent fortement la liste des groupes de Lie compatibles. A titre d’exemple et pour bien comprendre comment peut changer la physique quand on change la sym trie du MS nous allons tudier certains aspects du groupe $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ puis plus bri vement ceux du groupe $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{O}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. ##### Le groupe $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ Tout comme dans le Mod le Standard, il nous faut "ranger" nos particules dans les "casiers" disponibles de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Dans $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ nous disposons <sup>7</sup><sup>7</sup>7Une fa on simple de les trouver et de ”jouer” avec les tableaux d’Young. d’une repr sentation spinorielle fondamentale de dimension 5 et une repr sentation spinorielle antisym trique de dimension 10 pour les fermions d’une g n ration. Mais pour r partir ces quinze fermions ($`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\times }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\times }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{12}$}`$ quarks + 2 leptons charg s + 1 neutrino), il y a quelques r gles respecter. Tout d’abord, nous pouvons exprimer les repr sentations de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ en terme des repr sentations de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ : $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{𝟓}}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathbf{3,1}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{𝟏}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{𝟐}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ (4.11) $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{𝟓}}$}}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{𝟑}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathbf{,1}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathbf{1,2}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ (4.12) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{𝟏𝟎}}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{𝟑}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathbf{,1}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathbf{3,2}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathbf{1,1}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}`$ (4.13) Par exemple (4.11) veut dire que dans la repr sentation 5 de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, nous devons mettre un triplet de couleur, singulet de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, et un singulet de couleur, anti-doublet de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. De plus, il faut que la somme des charges dans chacun de ces 2 multiplets soit nulle<sup>8</sup><sup>8</sup>8La charge lectrique est une combinaison lin aire des g n rateurs diagonaux (dans le MS, on a $`\colorbox[rgb]{1,1,1}{$Q$}\colorbox[rgb]{1,1,1}{$=$}\colorbox[rgb]{1,1,1}{$T$}_\colorbox[rgb]{1,1,1}{$3$}\colorbox[rgb]{1,1,1}{$+$}\colorbox[rgb]{1,1,1}{$Y$}`$), qui sont de trace nulle, donc est de trace nulle aussi.. La seule combinaison possible qui soit en accord avec le MS <sup>9</sup><sup>9</sup>9Le choix des signes est fait en accord avec la d finition de la conjugaison de charge et de fa on obtenir les bonnes interactions du MS. est : $$\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{𝟓}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\begin{array}{c}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\\ \overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\\ \overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (4.14) et donc pour le reste des fermions de la première génération: $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{𝟏𝟎}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\sqrt{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\begin{array}{ccccc}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}& \overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}& \overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (4.15) où nous n gligeons les ventuels m langes entre les familles. Comme une famille enti re gauche entre dans ($`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{𝟏𝟎}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{𝟓}}$}}`$), nous devons donc r pliquer la classification pr c dente pour les 3 familles et en cons quence nous ne pouvons pas donner d’explication de la pr sence de 3 familles et de leurs diff rences dans $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Nous venons de parler des fermions, nous allons maintenant aborder les bosons de jauge. Les groupes de type $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ ont $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ g n rateurs de sym tries donc $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ a 24 bosons de jauge ($`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}`$ a 8 gluons, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ a 2 $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ et un $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}`$ etc). Parmis ces 24 bosons de jauge, 12 correspondent au MS et 12 sont nouveaux. Ils appartiennent toujours la repr sentation adjointe du groupe de jauge, et celle-ci est de dimension 24 pour $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Si comme pour les fermions nous d composons la repr sentation adjointe en repr sentations de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, et que nous identifions les bosons d j connus nous obtenons : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{𝟐𝟒}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{o}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{o}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{o}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}}{\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathbf{3,2}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{𝟑}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathbf{,2}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{o}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}}{\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathbf{8,1}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{,0}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}}{\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathbf{1,3}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{,0}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}}{\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathbf{1,1}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{,0}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (4.16) Le 3 me chiffre dans les parenth ses est l’hypercharge du sous-multiplet. Les nouveaux bosons sont appel s les leptoquarks $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{X}$}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}`$, sont charg s lectriquement <sup>10</sup><sup>10</sup>10Respectivement 4/3 et 2/3., sont color s et ont un isospin 1/2. Il peut donc y avoir des interactions directes entre quarks et leptons, nous en verrons les cons quences juste apr s. En notation matricielle, $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{24}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\begin{array}{ccccc}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}& \overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{X}$}}& \overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}& \overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{X}$}}& \overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}& \overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{X}$}}& \overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{X}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{X}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{X}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (4.17) o les $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}`$ sont les g n rateurs de SU(5) repr sent s par des matrices $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\times }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}`$ (l’ quivalent pour SU(5) des matrices de Pauli de SU(2)). Le choix de la base tant libre, nous avons choisi de les crire telles que $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}`$ agisse sur les 3 premi res lignes et colonnes et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}`$ sur les 2 derni res. D’o les deux blocs avec les gluons et les W. Le boson $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}`$ se situe sur la diagonale. Puisque $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ n’est qu’une sym trie valable tr s haute nergie, elle se brise plus basse nergie pour donner le MS. De plus, les multiplets fermioniques tant ceux de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ nous devons adapter le secteur de Higgs et trouver les bons multiplets qui brisent la sym trie lectrofaible. Pour briser $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ nous pouvons choisir (c’est le choix le plus simple) un multiplet adjoint $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Phi }}$}`$, de dimension 24, de bosons de Higgs. La v.e.v de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Phi }}$}`$ qui pr serve la couleur (3 premi res lignes et colonnes), l’isospin (2 derni res) et l’hypercharge (la diagonale) s’ crit : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Phi }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Phi }}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\begin{array}{ccccc}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}& \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (4.18) Cette premi re tape rend les bosons $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{X}$}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}`$ massifs en laissant les bosons du MS de masses nulles. $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{X}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{25}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{8}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Phi }}$}$$ (4.19) La v.e.v $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Phi }}$}`$ doit donc tre de l’ordre de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{15}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{16}$}}`$ GeV ($`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}}`$). La brisure lectrofaible se fait dans $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ de fa on minimale par un multiplet 5 qui contient un triplet de Higgs color s et le doublet de Higgs du MS : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{[}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{𝟓}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{]}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝐇}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝐂}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝐇}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (4.20) La v.e.v la plus simple qui brise seulement $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}`$ est de la forme : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\begin{array}{c}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\\ \colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (4.21) Les bosons lectrofaibles sont alors rendus massifs et de l’ordre de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$, $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (4.22) et les fermions acqui rent une masse de l’ordre de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$ gr ce aux couplages de Yukawa. Le potentiel effectif est donc fonction des champs $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Phi }}$}`$. En toute rigueur, outre une partie responsable de la premi re tape et une autre partie responsable de la brisure lectrofaible (de la m me forme que le potentiel du MS), nous devons introduire un terme qui m lange les deux champs. Le traitement des brisures est donc un peu plus complexe. Mais puisque les 2 chelles de brisure sont s par es par $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{13}$}}`$ GeV, elle peuvent tre consid r es raisonnablement ind pendantes. Cependant, cette hi rachie entre les masses des Higgs requiert un ajustement extr mement fin des param tres du potentiel, de 13 ordres de grandeurs, ce qui para t peu naturel. En effet, comme nous l’avons discut au premier chapitre, la masse du Higgs $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$, va recevoir des corrections venant des boucles de particules de masse $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}}`$. Pour que $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$ reste de l’ordre de 100 GeV, il faut des annulations accidentelles extr mement pr cises entre des contributions de l’ordre de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}}`$. Ce probl me peut tre r solu par l’apport de la supersym trie qui justifie ces annulations. Autre nouveaut de ce secteur de Higgs : l’existence de Higgs color s. Leurs changes conduisent de nouvelles interactions qui n’existaient pas dans le MS qui peuvent se r v ler tr s importantes (voir plus loin). De plus, d’autres multiplets de Higgs peuvent exister si nous voulons briser $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ en $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}`$. Le secteur de Higgs peut tre beaucoup plus compliqu dans les GUTs que dans le MS et peut faire intervenir des repr sentations plus grandes. ##### Une autre possibilit : $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{O}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ Une autre possibilit tr s souvent tudi e est le groupe d’unification $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{O}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. C’est un groupe de rang 5, $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{O}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (4.23) L’avantage principal de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{O}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ sur $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ est qu’il poss de une repr sentation spinorielle fondamentale de dimension 16 ce qui permet de ranger tous les fermions d’une g n ration et un neutrino droit, en une seule fois . Ceci en fait un groupe "naturel" du point de vue de la physique des neutrinos et du m canisme de Seesaw. Dans $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, les neutrinos droits sont rajouter " la main" comme singulets. Mais la structure des interactions de jauge $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{O}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ n’a aucune incidence sur les neutrinos, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{O}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ est juste plus adapt si on croit au Seesaw. En termes des repr sentations de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{𝟏𝟔}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{𝟏𝟎}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{𝟓}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{𝟏}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (4.24) Dans $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{O}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$, le nombre de bosons de jauge s’el vent 45, ce qui nous fait 33 bosons suppl mentaires par rapport au MS et donc beaucoup d’interactions possibles. De plus, la brisure de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{O}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ est plus complexe car se fait en deux tapes : en passant par $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ ou par $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$ puis intervient la brisure de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Le secteur de Higgs est tr s tendu et fait intervenir de larges multiplets, de dimension 10, 16, 45, 120, 126 selon le mod le. C’est un secteur extr mement important pour la ph nom nologie. ##### Le mod le minimal $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ supersym trique Une fois la supersym trie introduite pour r soudre les probl mes de hi rarchie des chelles (chapitres 1 et 2), nous avons vu au d but de ce chapitre qu’elle permettait en plus une unification beaucoup plus pr cise des couplages de jauge, la rendant presque indispensable pour les mod les Grand-Unifi s. Dans un mod le minimal $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ supersym trique, les multiplets vont tre remplac s par des supermultiplets. Les repr sentations utilis es sont ($`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{𝟏𝟎}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{𝟓}}$}}`$) pour les champs de mati re et leurs sfermions, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{𝟐𝟒}}$}`$ pour les bosons et leurs partenaires fermioniques les "bosinos". Pour les m me raisons qu’au chapitre 2, nous devons introduire 2 supermultiplets, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{𝟓}}$}`$ et $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{𝟓}}$}}`$, pour les Higgs et les Higgsinos, et enfin un supermultiplet <sup>11</sup><sup>11</sup>11Il n’y en a qu’un seul, car il est auto-conjugu . $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{𝟐𝟒}}$}`$ de Higgs (et Higgsinos) pour briser $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. L’existence de scalaires suppl mentaires permet d’autres interactions et notamment d’autres modes de d sint gration pour le proton (via change de Higgsinos par exemple, ou de squarks si la $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$-parit n’est pas conserv e) #### 4.1.3 Les cons quences typiques des GUTs ##### La d sint gration du proton Une cons quence directe de l’existence d’une Grande Unification est la d sint gration du proton. En effet, les bosons $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{X}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}`$ et les triplets de Higgs couplent les indices de couleurs de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ avec les indices d’isospin faible de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Nous avons donc des interactions possibles du genre $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{X}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}`$,…, comme montr par les diagrammes de Feynman de la figure (4.3). Autrement dit, dans les mod les GU il y a des interactions telles que $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Delta }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$, le nombre baryonique n’est pas conserv . Rien n’emp che alors le proton de se d sint grer. Les modes de d sint gration possibles incluent: $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{K}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{K}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}$$ (4.25) Pour valuer le temps de vie du proton nous pouvons consid rer une interaction effective basse nergie entre 4 fermions (qui repr sente par exemple un processus baryon$`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ lepton + m son). L’interaction est m di e par l’ change d’un boson $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{X}$}`$ ou $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}`$ tr s lourd (la GU est suppos e se r aliser $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{16}$}}`$ GeV et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{X}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒪}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$) donc l’interaction peut tre consid r e comme ponctuelle. L’amplitude est proportionnelle <sup>12</sup><sup>12</sup>12Par analogie avec la constante de Fermi des interactions faibles, $`\colorbox[rgb]{1,1,1}{$G$}_\colorbox[rgb]{1,1,1}{$F$}\colorbox[rgb]{1,1,1}{$=$}\frac{\colorbox[rgb]{1,1,1}{$g$}^\colorbox[rgb]{1,1,1}{$2$}}{\colorbox[rgb]{1,1,1}{$8$}\colorbox[rgb]{1,1,1}{$m$}_\colorbox[rgb]{1,1,1}{$W$}^\colorbox[rgb]{1,1,1}{$2$}}`$. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{X}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{X}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{8}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{X}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}}`$, et prend la forme suivante dans le mod le $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ minimal: $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{X}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{X}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}`$ (4.26) où nous avons explicit les chiralit s des fermions et nous n gligeons les ventuels m langes entre les diff rentes g n rations de quarks et leptons. Une interaction de la forme (4.26) donnerait la forme suivante du taux de d sint gration: $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Gamma }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{X}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (4.27) Le facteur $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}`$ ($`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}`$ est la masse du proton) vient de consid rations dimensionnelles <sup>13</sup><sup>13</sup>13Nous rappelons qu’une largeur de d sint gration $`\colorbox[rgb]{1,1,1}{$\mathrm{\Gamma }$}`$ a la dimension d’une masse. et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}`$ est un facteur qui d pend du mod le de la Grande Unification et de la dynamique hadronique du proton (qui est un ensemble complexe de 3 quarks confin s, ce qui am ne des complications au calcul). Les calculs d taill s donnent dans le cas $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ non-supersym trique un temps de vie du proton entre $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2.5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\times }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{28}$}}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1.6}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\times }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{30}$}}`$ ans . Ceci est en d saccord avec les mesures de SuperKamiokande: $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1.6}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\times }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{33}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{ans}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (4.28) De ce point de vue l aussi les pr dictions du mod le non-supersym trique ne sont donc pas en accord avec les donn es exp rimentales. En revanche, les mod les supersym triques ne sont g n ralement pas exclus par cette limite sur la d sint gration du proton, et ceci est notamment d une chelle d’unification plus haute (donc des masses des $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{X}$}`$ ou $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Y}$}`$ plus lev es). Typiquement, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{35}$}}`$ ans dans les mod les supersym triques. Les futurs projets exp rimentaux (Fr jus, Hyperkamiokande,…) devraient atteindre cette limite. L’observation de la d sint gration du proton tant un des tests les plus importants pour les GUTs, elle est de ce fait tr s attendue. ##### Les masses des quarks et leptons Les masses des quarks et leptons sont donn es, comme dans le MS, la brisure de la sym trie lectrofaible $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Quarks et leptons se situent dans les m mes repr sentations ce qui implique des relations particuli res entre leurs masses. Dans $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ par exemple, on a les multiplets de fermions $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{𝟓}}$}}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{𝟏𝟎}}$}`$ qui contiennent respectivement $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}`$, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}`$ (pour la premi re famille). Un terme de masse est donc proportionnel <sup>14</sup><sup>14</sup>14Nous rappelons que $`\overline{\colorbox[rgb]{1,1,1}{$\psi $}}\colorbox[rgb]{1,1,1}{$\psi $}\colorbox[rgb]{1,1,1}{$=$}\overline{\colorbox[rgb]{1,1,1}{$\psi $}}_\colorbox[rgb]{1,1,1}{$R$}\colorbox[rgb]{1,1,1}{$\psi $}_\colorbox[rgb]{1,1,1}{$L$}\colorbox[rgb]{1,1,1}{$+$}\overline{\colorbox[rgb]{1,1,1}{$\psi $}}_\colorbox[rgb]{1,1,1}{$L$}\colorbox[rgb]{1,1,1}{$\psi $}_\colorbox[rgb]{1,1,1}{$R$}\colorbox[rgb]{1,1,1}{$=$}\colorbox[rgb]{1,1,1}{$$}\colorbox[rgb]{1,1,1}{$\psi $}_\colorbox[rgb]{1,1,1}{$L$}^\colorbox[rgb]{1,1,1}{$\mathrm{T}$}\colorbox[rgb]{1,1,1}{$C$}\colorbox[rgb]{1,1,1}{$\psi $}_\colorbox[rgb]{1,1,1}{$L$}^\colorbox[rgb]{1,1,1}{$c$}\colorbox[rgb]{1,1,1}{$+$}\colorbox[rgb]{1,1,1}{$h$}\colorbox[rgb]{1,1,1}{$.$}\colorbox[rgb]{1,1,1}{$c$}\colorbox[rgb]{1,1,1}{$.$}`$. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{𝟓}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{𝟏𝟎}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ et pour former un scalaire on peut coupler ce terme un Higgs dans une repr sentation $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{𝟓}}$}`$ ce qui nous donne : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒴}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{[}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{𝟓}}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{]}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{T}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{[}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{𝟏𝟎}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{]}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{[}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{𝟓}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{]}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (4.29) Apr s la brisure, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{[}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{𝟓}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{]}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{[}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{𝟓}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{]}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}`$, on a le terme de masse $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒴}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\overset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (4.30) La masse de l’ lectron est donc la m me que celle du quark down, ainsi que pour les deux autres familles, l’ chelle $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}}`$ : $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}`$ (4.31) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}`$ (4.32) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}`$ (4.33) Mais il nous faut encore renormaliser jusqu’ basse nergie ou nous connaissons la valeur des masses. Pour la troisi me famille, nous avons approximativement : $$\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left[}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right]}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{7}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (4.34) ce qui est ( peu pr s) observ exp rimentalement <sup>15</sup><sup>15</sup>15Pour $`\colorbox[rgb]{1,1,1}{$m$}_\colorbox[rgb]{1,1,1}{$b$}\colorbox[rgb]{1,1,1}{$$}\colorbox[rgb]{1,1,1}{$4.1$}\colorbox[rgb]{1,1,1}{$$}\colorbox[rgb]{1,1,1}{$4.9$}`$ GeV et $`\colorbox[rgb]{1,1,1}{$m$}_\colorbox[rgb]{1,1,1}{$\tau $}\colorbox[rgb]{1,1,1}{$$}\colorbox[rgb]{1,1,1}{$1.78$}`$ GeV nous trouvons que le rapport exp rimental est autour de 2.5. .. En incluant les corrections radiatives d’ordres sup rieurs, la supersym trie, etc, il y a toujours accord entre la relation pr dite et l’exp rience. Malheureusement, la renormalisation appliqu e aux 2 premi res familles donnent des r sultats en contraction avec les mesures. Ceci veut dire qu’il faut construire un mod le non-minimal et y inclure d’autres multiplets de Higgs (des 45 par exemple). Mais d’une fa on g n rale, l’organisation nouvelle des particules propos e par les mod les Grand-Unifi s implique des relations particuli res entre les masses des quarks et leptons. Beaucoup de travaux ont donc pour but d’expliquer dans ce type de mod le le spectre de masses observ . #### 4.1.4 Discussion R sumons cette section sur les mod les Grand-Unifi s en passant en revue les divers avantages et faiblesses de ce type de mod les. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ L’organisation des fermions d’une famille dans les repr sentations plus grandes ($`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{𝟏𝟎}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{𝟓}}$}}`$ dans le cas de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$) rend la structure des quarks et leptons beaucoup plus simple que dans le MS. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ Les quarks et leptons sont dans les m mes multiplets ce qui explique la charge lectrique fractionnaire (multiple de 1/3) des quarks et donc la quantification de la charge. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ Les masses des quarks et leptons sont reli es et certains rapports pr dits (par exemple $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{b}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\tau }$}`$ Cependant, une relation viable pour les deux premi res g n rations demanderait, par exemple, un secteur de Higgs non-minimal ou des op rateurs de dimensions 5. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ Ces mod les r alisent pleinement l’unification des interactions (une seule constante de couplage pour la partie non-gravitationnelle) et haute nergie ce qui est en accord avec la stabilit observ du proton. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ Ces mod les sont aussi justifi s par l’unification pr cise des couplages sugg r e exp rimentalement (c’est- -dire la pr diction pr cise de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{sin}}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\theta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$ ou de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}`$ partir d’une unification postul e) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ Malheureusement, en g n ral les GUTs ne donnent pas d’explication sur l’origine du nombre de familles ni sur l’origine de la saveur (matrices CKM et MNS,…) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ Une unification si haute nergie implique un grand "d sert" de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{12}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{13}$}}`$ GeV. Qu’y a t’il vraiment dans cette gamme d’ nergie, est-ce vraiment "d sert" ? $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ La plupart des GUT ph nologiquement viables ont un secteur de Higgs compliqu incluant beaucoup de multiplets et de repr sentations de dimensions lev es. Les mod les Grand Unifi s tels que pr sent s ici n’ont de toute fa on pas la pr tention de tout expliquer mais il semble tr s int ressant et attirant d’inclure cette id e et la plupart des ingr dients dans des mod les r alistes d crivant la physique jusqu’ de tr s hautes chelles d’ nergie. Pour tester ces id es, il faudrait en particulier pouvoir observer la d sint gration du proton (qui contraint en particulier le secteur de Higgs) et la supersym trie qui est quasi-indispensable aux GUTs, et d’enrichir notre connaissance de la physique du neutrino (qui en quelque sorte "sonde" des nergies proches des GUTs). ### 4.2 La brisure de la supersym trie La brisure de la supersym trie est n cessaire. En effet, nous n’observons pas de particules supersym triques de m me masses que leurs partenaires usuels: $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}}}`$. Or, l’alg bre de supersym trie implique que deux tats li s par une transformation de supersym trie ont les m mes valeurs propres pour l’op rateur $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ donc la m me masse. Il faut donc briser la supersym trie. La question est de savoir si la brisure est explicite c’est- -dire pr sente dans le Lagrangien sous-jacent la th orie ou bien, si la brisure est spontan e, c’est- -dire induite par un tat de vide non supersym trique. Plusieurs raisons jouent en d faveur d’une brisure explicite. C’est d’une part non esth tique, ensuite ce n’est pas une mani re analogue celle avec laquelle sont bris es les sym tries de jauge, enfin, cela conduirait des inconsistances dans les th ories de supergravit . Ainsi, les th oriciens se sont concentr s sur la brisure spontan e de la supersym trie. Si le vide est non supersym trique, il existe un tat fermionique $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}`$ qui est coupl au vide par l’interm diaire de l’op rateur fermionique $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}`$ (qui correspond la charge supersym trique) : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\hspace{0.33em}0}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (4.35) Le fermion $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}`$ est l’ quivalent du boson de Goldstone dans les symmetries bosoniques spontan ment bris es. Le champ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}`$ tant un champ fermionique, on l’appelle donc fermion de Goldstone ou Goldstino. Jusqu’ pr sent, nous n’avons parl de supersym trie que dans le sens d’une sym trie globale, c’est- -dire dont les transformations ne d pendent pas de l’espace-temps, donc n’incluant pas la gravit . C’est dans ce cadre que l’on se place pour le moment : ainsi ce que l’on brise ce stade est une sym trie globale. Or, un probl me appara t lorsqu’on brise la supersym trie globale: l’ nergie du vide est positive. Pour le constater, il suffit de voir la valeur dans le vide de l’anticommutateur des charges $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}`$: $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (4.36) et donc en introduisant la relation de fermeture $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (4.37) nous constatons ainsi que la brisure de la supersym trie globale entra ne: $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\chi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\hspace{0.33em}0}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (4.38) On obtient une valeur de l’ nergie du vide strictement positive. #### 4.2.1 Terme F et terme D Afin de voir comment on peut obtenir une valeur de l’ nergie du vide non nulle, reprenons le potentiel effectif d’une th orie supersym trique globale: $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Sigma }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Sigma }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (4.39) Rappelons que le premier terme est appel terme F, et le second est le terme D. Ainsi pour obtenir une valeur de l’ nergie du vide non nulle, il faut soit que le terme F soit d fini positif (valeur moyenne dans le vide non nulle) soit que le terme D soit d fini positif. ##### Terme D L’option $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{D}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ implique de construire un mod le avec un groupe de sym trie de jauge $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. L’exemple le plus simple consiste prendre un supermultiplet chiral avec une charge unit pour lequel le potentiel effectif est: $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{D}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\xi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (4.40) Le terme suppl mentaire $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\xi }$}`$ n’est pas permis dans une th orie non-ab lienne c’est la raison pour laquelle il nous faut utiliser une th orie qui poss de un groupe de jauge $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Le minimum du potentiel effectif est atteint pour $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ et alors $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{D}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\xi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ et la supersym trie est spontan ment bris e. Dans cet tat de vide, on trouve que: $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\xi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\hspace{0.33em}0}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\hspace{0.33em}0}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (4.41) On distingue nettement dans cet exemple la diff rence de masse entre boson et fermion qui correspondent au supermultiplet $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Malheureusement, on ne peut pas utiliser le groupe de jauge $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ de l’ lectromagn tisme dans le mod le standard. En effet, dans le Mod le Standard, il y a des champs qui ont des signes diff rents pour l’hypercharge permettant au potentiel effectif de s’annuler. On doit rajouter un nouveau groupe de jauge $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ mais aussi de nombreux nouveaux champs pour annuler les anomalies triangulaires suppl mentaires qui apparaissent cause de ce nouveau groupe de jauge. Ainsi, le mod le de brisure de supersym trie avec un terme D n’a pas suscit norm ment d’int r t bien qu’aujourd’hui il connaisse un renouveau dans le cadre des mod les d riv s des th ories de cordes. ##### Terme F L’option $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ implique des champs chiraux suppl mentaires avec des couplages artificiels: ceux du Mod le Standard ne suffisent pas; l’exemple le plus simple consiste prendre trois supermultiplets chiraux $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}`$ avec le superpotentiel suivant: $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (4.42) On trouve les termes F correspondants : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\hspace{0.33em}2}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (4.43) et le potentiel effectif se r crit: $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Sigma }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\hspace{0.33em}4}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\beta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (4.44) On v rifie que les trois termes de l’ quation pr c dente ne peuvent pas s’annuler simultan ment. Ainsi, on obtient n cessairement $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ et la supersym trie est bris e. Pour conclure cette section, nous pouvons dire qu’il n’existe pas de moyen satisfaisant de briser la supersym trie globale car d’une fa on ou d’une autre, on est amen soit rajouter de nouveaux champs de mati re et des couplages artificielles, soit introduire un nouveau groupe de jauge et d’autres champs. Dans les ann es $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1980}$}`$, des tentatives ont t faites pour "dissimuler" les champs suppl mentaires dans un secteur cach mais il s’est r v l qu’il tait tr s compliqu d’obtenir un mod le ph nom nologiquement viable. De plus, la brisure de la supersym trie globale implique n cessairement une nergie du vide non nulle: premi re vue, cela n’est pas une mauvaise chose; en effet, les observations actuelles tendent prouver que l’ nergie du vide n’est pas nulle. Le probl me fondamental vient du fait que la mesure de cette nergie du vide est compl tement en d saccord avec la pr diction de l’ nergie du vide de la supersym trie globale: la valeur mesur e de l’ nergie du vide est: $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{123}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (4.45) alors que l’ nergie du vide de la supersym trie brisée pr dit une constante cosmologique: $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{64}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (4.46) Nous obtenons une diff rence de 60 ordres de grandeur! Pour discuter de la brisure de la supersym trie de mani re plus satisfaisante, il faut une th orie supersym trique incluant la gravit : c’est l’objectif du prochain chapitre. #### 4.2.2 Supersym trie locale ou Supergravit Jusqu’ pr sent, nous avons consid r les transformations globales de supersym trie dans lesquelles le spineur des transformations infinit simales $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}`$ est ind pendant de l’espace-temps. Nous allons maintenant consid rer un spineur de transformations d pendant de l’espace-temps $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ par analogie avec les symm tries bosoniques qui, une fois rendu locales, donnent naissance aux th ories de jauge. De plus, le fait de rendre la supersym trie locale conduit un m canisme analogue au m canisme de Higgs pour briser les sym tries bosoniques: le m canisme de super-Higgs qui permettra de briser de mani re l gante la supersym trie. En prime, une th orie locale de la supersym trie contient n cessairement, comme nous le verrons, la gravit (c’est la raison pour laquelle cette th orie se nomme la Supergravit ) et ouvre la perspective d’unifier toutes les interactions des particules et les champs de mati re avec des transformations de supersym trie tendue: $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{q}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{J}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (4.47) o $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}`$ est le graviton de spin 2 et $`\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}}`$ est le gravitino, son partenaire supersym trique de spin-3/2, qui s’inscrivent tous deux dans le supermultiplet de la gravitation: $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\begin{array}{c}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}\\ \stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}}\end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\begin{array}{c}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\\ \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\end{array}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (4.48) La Supergravit est un ingr dient essentiel pour la compr hension des interactions gravitationnelles des particules supersym triques, et donc n cessaire une discussion coh rente de l’énergie du vide. #### 4.2.3 Supergravit Pour comprendre pourquoi le fait de rendre la supersym trie locale implique une pr sence de la gravit dans la th orie, consid rons ce qui arrive si l’on applique un supermultiplet chiral les deux transformations de supersym trie suivantes: $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}`$ $`\sqrt{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}`$ (4.49) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\sqrt{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}`$ (4.50) Si l’on construit le commutateur sur les champs $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}`$ ou $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}`$, on obtient: $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{[}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{]}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (4.51) On observe que l’effet sur les champs est quivalent une translation d’espace-temps, puisque $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}`$. Si les transformations spinorielles infinit simales $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}`$ sont ind pendantes de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}`$, la translation est globale et la th orie est invariante par translation globale. Mais si les $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}}`$ d pendent de la position dans l’espace-temps, on obtient que le commutateur appliqu au champ est quivalent un changement de coordonn es locales sur les champs et donc la th orie est invariante par changement de coordonn es locales. Or, nous savons qu’une th orie invariante par changement de coordonn es locales contient n cessairement la gravit : l’invariance par changement de coordonn es locales est un des points de d part de la construction de la Relativit G n rale qui est la th orie de la gravit . ##### Analogie avec les th ories de jauge Dans cette section nous verrons que le gravitino merge naturellement comme champ de jauge de la supersym trie. Consid rons, dans un premier temps, la variation du terme cin tique d’un fermion sous une transformation de jauge : $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Dans une transformation de jauge, le champ fermionique devient : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (4.52) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ est une variation de phase d pendante de la position dans l’espace-temps. Ainsi, un terme suppl mentaire, par rapport une th orie o la phase est ind pendante de l’espace-temps, appara t dans la variation du terme cin tique: $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (4.53) Pour annuler cette variation et maintenir l’invariance du lagrangien sous les transformations de jauge, on introduit un champ dit champ de jauge: $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Dans une sym trie de jauge ab lienne, il intervient dans le Lagrangien sous cette forme: $$\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}$$ (4.54) et se transforme ainsi : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (4.55) Dans le cas de transformations locales de supersym trie, la variation du champ fermionique est: $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (4.56) et ainsi, la variation du terme cin tique fermionique contient un terme: $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}$$ (4.57) qui est compens par l’introduction d’un terme contenant un nouveau champ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}`$ : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\kappa }$}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (4.58) qui se transforme comme ceci sous la supersym trie: $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\kappa }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (4.59) Le nouveau champ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}`$ est un spineur avec un indice de Lorentz : donc il possède spin 3/2. Il est le champ de jauge de la supersym trie : c’est un fermion de jauge. Comme il a un spin 3/2, il ne peut tre que le partenaire supersym trique du graviton de spin 2: c’est donc le gravitino. ##### Le Lagrangien de pure Supergravit Consid rons le plus simple Lagrangien pour un gravitino et un graviton. Il consiste en la somme du Lagrangien de Einstein-Hilbert pour la Relativit G n rale (description du graviton) et celui de Rarita-Schwinger pour un champ de spin 3/2 (descrption du gravitino). Bien entendu, il faut rendre invariant le Lagrangien de Rarita-Schwinger sous transformations g n rales de coordonn es en covariantisant la d riv e: $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\kappa }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\sqrt{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}}\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒟}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\sigma }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (4.60) o $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}`$ est le tenseur m trique: $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\eta }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}$$ (4.61) o $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}`$ est le vierbein (qui d crit, entre autres, le champ du graviton) et $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒟}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\omega }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{[}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{]}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (4.62) est la d riv e covariante avec $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\omega }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}`$ la connexion de spin. Ce Lagrangien est naturellement invariant sous transformations locales de supersym trie. Les champs se transforment de la mani re suivante: $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{ϵ}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}`$ $`\overline{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\gamma }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}`$ (4.63) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\omega }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\hspace{0.33em}0}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}`$ (4.64) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\delta }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}`$ $`{\displaystyle \frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\kappa }$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒟}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}`$ (4.65) #### 4.2.4 Le m canisme de Super-Higgs Dans le Lagrangien pr c dent, nous n’avons pas inclus le couplage la mati re, dont nous parlerons plus tard. Cependant, nous pouvons d j parler du m canisme de brisure spontan e de la supersym trie locale: le m canisme de super-Higgs. En effet, nous venons d’introduire le gravitino qui se trouve au centre de ce m canisme. Rappelons que dans le m canisme de Higgs conventionnel, un boson de Goldstone sans masse (de spin 0 donc avec une seule polarisation) est absorb par un boson de jauge sans masse (d’h licit + ou -1 donc avec deux états de polarisation) qui va acqu rir par ce processus les degr s de libert du boson de Goldstone et donc poss der trois tats de polarisations qui lui permettent d’acqu rir une masse: $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\times }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\hspace{0.33em}1}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\times }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\hspace{0.33em}3}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\times }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (4.66) Dans une th orie localement supersym trique, les deux tats de polarisation (h licit + ou - 1/2) du fermion de Goldstone sans masse sont absorb s par un gravitino sans masse d’h licit + ou - 3/2, donc avec deux tats de polarisation, pour lui donner les quatres tats de polarisations n cessaire l’obtention d’une masse: $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\times }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\hspace{0.33em}2}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\times }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{F}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\hspace{0.33em}4}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\times }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\psi }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (4.67) Ce processus brise la supersym trie locale, puisqu’il donne une masse au gravitino alors que le graviton n’en a pas: $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$. C’est le seul moyen consistent pour briser la supersym trie. De plus, il peut se faire en donnant une nergie nulle au vide: $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\hspace{0.33em}0}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\hspace{0.33em}0}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (4.68) Ainsi, on peut obtenir la fois une brisure de la supersym trie locale et une constante cosmologique nulle, ce qui n’ tait pas le cas de la brisure globale de supersym trie. #### 4.2.5 Couplage de la Supergravit la mati re Le Lagrangien complet de la Supergravit incluant les multiplets vecteurs et chiraux peut tre obtenu par la m thode de calcul tensoriel locale. C’est un travail laborieux. Nous ne donnerons que quelques l ments cl s sans crire de preuves. Parmi ces expressions nouvelles par rapport la supersym trie globale, on trouve principalement le potentiel de K hler. Le potentiel de K hler est une fonction des champs scalaires (contenus dans les supermultiplets chiraux) : $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Le superpotentiel est reli au potentiel de K hler. Ce dernier est aussi appel vari t de K hler, car il d crit une g om trie interne qui influe directement sur les termes cin tiques des champs fermioniques et scalaires et qui d termine la masse des particules apr s la brisure de supersym trie. C’est aussi un param tre d’ordre pour la brisure de supersym trie: $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^{\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (4.69) o $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ est la valeur dans le vide de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}`$ est la valeur dans le vide du superpotentiel $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}`$<sup>16</sup><sup>16</sup>16O $`\colorbox[rgb]{1,1,1}{$W$}\colorbox[rgb]{1,1,1}{$=$}\colorbox[rgb]{1,1,1}{$W$}_{\colorbox[rgb]{1,1,1}{$c$}\colorbox[rgb]{1,1,1}{$a$}\colorbox[rgb]{1,1,1}{$c$}\colorbox[rgb]{1,1,1}{$h$}}\colorbox[rgb]{1,1,1}{$+$}\colorbox[rgb]{1,1,1}{$W$}_{\colorbox[rgb]{1,1,1}{$o$}\colorbox[rgb]{1,1,1}{$b$}\colorbox[rgb]{1,1,1}{$s$}}`$ et $`\colorbox[rgb]{1,1,1}{$<$}\colorbox[rgb]{1,1,1}{$W$}_{\colorbox[rgb]{1,1,1}{$o$}\colorbox[rgb]{1,1,1}{$b$}\colorbox[rgb]{1,1,1}{$s$}}\colorbox[rgb]{1,1,1}{$>$}\colorbox[rgb]{1,1,1}{$=$}\colorbox[rgb]{1,1,1}{$0$}`$, mais $`\colorbox[rgb]{1,1,1}{$<$}\colorbox[rgb]{1,1,1}{$W$}_{\colorbox[rgb]{1,1,1}{$c$}\colorbox[rgb]{1,1,1}{$a$}\colorbox[rgb]{1,1,1}{$c$}\colorbox[rgb]{1,1,1}{$h$}}\colorbox[rgb]{1,1,1}{$>$}\colorbox[rgb]{1,1,1}{$$}\colorbox[rgb]{1,1,1}{$0$}`$.. Comme nous l’avons dit, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}`$ d termine les termes cin tiques des champs fermioniques et des champs scalaires. Dans le cas des champs scalaires: $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{K}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (4.70) o $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}`$ est la m trique de K hler: $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (4.71) Le potentiel effectif sans la pr sence des termes $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{D}$}`$ est aussi d termin par le potentiel de K hler: $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{[}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{j}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{]}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{v}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (4.72) On remarque que ce potentiel peut s’annuler sans emp cher la brisure de supersym trie, ce qui n’ tait pas le cas dans la supersym trie globale. En effet, on peut avoir simultan ment $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$. L’ tat de vide de la th orie doit correspondre un minimum de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}`$. Or, il se trouve que pour des formes g n rales de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}`$, et pour certaines valeurs des champs, le potentiel $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}`$ devient n gatif: ceci constitue une catastrophe pour la cosmologie, puisque cela signifierait que la constante cosmologique est n gative et donc que l’Univers s’effondre sur lui-m me. Heureusement, il existe une classe particuli re de potentiels de K hler, non nǵatifs. Il se trouve que c’est la classe de potentiels qui merge des th ories de cordes. De m me qu’il existe un potentiel de K hler qui d termine la g om trie et la cin tique des champs chiraux, il existe une fonction appel e fonction cin tique not e $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ qui d termine les termes cin tiques des supermultiplets vecteurs. #### 4.2.6 Th orie effective aux basses nergies La th orie basse- nergie trouve son origine dans une th orie haute- nergie : la th orie la plus g n rale de supergravit n’est pas renormalisable, mais les termes non renormalisables ne sont pas importants à basses énergies. Donc, en première approximation on peut ne garder que des termes renormalisables dans le potentiel effectif de la supergravit basses nergies. Pour des choix g n riques du potentiel de K hler, on obtient des masses issues de la brisure de la supersym trie non nulles pour les gauginos: $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (4.73) L’universalit des masses des jauginos correspondant aux diff rents groupes de jauge $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ n’est pas syst matique, mais cela merge naturellement si la g om trie n’est pas trop compliqu e, par exemple si la fonction cin tique $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ est un singulet. En d veloppant le potentiel effectif, on trouve des termes proportionnels $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$, qui sont interpr t s comme les masses des scalaires issues de la brisure de la supersym trie: $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (4.74) Dans ce cas, il n’y a aucune raison th orique d’avoir universalit des masses: les mod les issus des th ories de cordes brisent souvent cette universalit . Par contre, il y a de bonnes raisons ph nom nologiques pour penser que les masses des scalaires avec la m me charge sont identiques. En effet, dans le cas contraire, on aurait des interactions v hicul es par des Sparticules virtuelles qui provoqueraient des changements de saveurs. Des formes g n riques du potentiel effectif conduisent aussi à la pr sence de termes d’interaction trilin aires issues de la brisure de supersym trie entre particules scalaires : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (4.75) Encore une fois, l’universalit n’a pas de raison fondamentale. Si la th orie supersym trique inclut aussi des termes d’interaction bilin aires comme c’est le cas dans l’extension minimale du Mod le Standard la supersym trie (MSSM) avec des termes $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ o le scalaire en question est le Higgs, on s’attend trouver aussi des termes $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ dans le potentiel effectif. La forme finale du Lagrangien de brisure ’explicite’ de la th orie basse nergie de la supergravit sugg r e par la brisure spontan e de la supersym trie locale est: $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Sigma }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\stackrel{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{~}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Sigma }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Sigma }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Sigma }}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{Herm}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{Conj}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (4.76) qui contient de nombreux param tres libres. La brisure de la supersym trie est explicite dans la th orie basse nergie mais, comme nous l’avons d j vu, elle est n anmoins douce car la renormalisation des param tres $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}`$ est logarithmique, sans corrections quadratiques. Il faut souligner que ces param tres ne sont pas fondamentaux, et que le m canisme de brisure de la supersym trie sous-jacent est spontan e. La renormalisation logarithmique des param tres signifie que l’on peut calculer leurs valeurs de basse- nergie partir de ses valeurs hautes nergies provenant d’une th orie des supercordes ou de supergravit . Pour le cas de la masse basse- nergie des gauginos $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}`$, on a : $$\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}}}$$ (4.77) une boucle, o $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}`$ est le couplage de jauge et o l’on a suppos l’unification des couplages l’ chelle de la supergravit . Pour le cas des scalaires, on a : $$\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{16}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pi }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{[}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{]}$}$$ (4.78) avec $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{ln}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ et o les coefficients li s au groupe de jauge ont t supprimés. Dans le cas des deux premi res g n rations, le premier terme dans la partie droite de l’ quation peut tre omis, car les couplages de Yukawa sont faibles. On obtient ainsi : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (4.79) dans le cas où les masses initiales sont identiques pour les scalaires, et les coefficients $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{C}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}`$ sont calculables dans tous les mod les. Le premier terme dans la partie droite de l’ quation (4.78) est important pour la troisi me g n ration et pour les bosons de Higgs du MSSM. En effet, le signe du premier terme est positif et celui du second est n gatif. Cela signifie que le dernier terme augmente $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ lorsque l’ chelle de renormalisation $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}`$ diminue, alors que le terme positif fait diminuer $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ lorsque $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Q}$}`$ diminue. Dans le cas du boson de Higgs $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}`$, le terme positif n’est pas n gligeable car $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}`$ a un couplage de Yukawa lev avec le quark top: $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{2,3}}$}}`$. Une observation tr s int ressante est que la brisure de sym trie spontan e lectrofaible est expliquable. En effet, on retrouve la forme du potentiel de brisure lectrofaible de mani re naturelle puisque $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{H}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{u}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ devient n gatif à basse énergie. Ainsi, la brisure spontan e de supersym trie entraine la brisure spontan e de la sym trie lectrofaible. Cela se produit une chelle d’ nergie exponentiellement plus petite que l’ chelle de la supergravit : $$\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{𝒪}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{:}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\alpha }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\lambda }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pi }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (4.80) Les calculs d’ volution permettent de démontrer que $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{W}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{100}$}`$ GeV merge naturellement si $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{60}$}`$ to $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{200}$}`$ GeV, ce qui a été confirm ensuite par des expériences. #### 4.2.7 Conclusion sur la brisure de la supersym trie locale Comme nous l’avons vu, la brisure de supersym trie locale serait induite par le m canisme de super-Higgs. Cependant, le d tail du m canisme de brisure reste un grand myst re. Pour que les sparticules soient plus lourdes que les particules, tout en restant une masse inf rieure 10 Tev (pour permettre la stabilisation des bosons de Higgs), une m thode consiste ajouter des champs non pr sents dans la th orie. On appelle ce secteur, le secteur cach . Dans certains mod les, le secteur cach communique gravitationnellement avec le secteur observable. Dans d’autres mod les, les messagers sont des bosons de jauge. Dans le secteur cach , on suppose l’existence d’une interaction forte qui permet la brisure dynamique de la supersym trie à une échelle élevée. Dans le cas de l’interaction forte usuelle, le passage de la zone perturbative la zone non pertubative de la th orie basse nergie provoque la condensation et le confinement des quarks. Par analogie, on attend un condensat de jauginos dans le secteur cach , ce qui briserait la supersym trie. Par exemple, dans la th orie des cordes h t rotiques dont le groupe de jauge est $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{8}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\times }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{8}$}`$, le premier groupe exceptionnel contient le mod le standard alors que le second facteur d crit le secteur cach , dans lequel apparait le condensat de jauginos. Les th ories de cordes donnent naissance des th ories spécifiques de supergravit dites no-scale. Dans ces mod les, la masse du gravitino est ind termin e au niveau de l’arbre car le potentiel effectif est constant. Dans un tel mod le, il n’y a qu’une seule échelle de masse au d part: celle de Planck. La masse du gravitino provient alors des corrections radiatives au potentiel effectif. La forme de ce potentiel au niveau de l’arbre ne fournit pas de termes de masse provenant de la brisure de supersym trie pour les scalaires. Pourtant, il existe des moyens de communiquer la brisure de la supersym trie au secteur observable. Par exemple, dans ces th ories de "no-scale" supergravit provenant des th ories de cordes, il existe souvent un second champ scalaire du secteur cach qui param trise la fonction cin tique des bosons de jauge et permet aux gauginos d’obtenir une masse. Les masses des particules scalaires sont alors donn es par les corrections radiatives lorsque l’on am ne les param tres de la th orie vers l’ chelle lectrofaible. Pour conclure ce paragraphe, il est int ressant de regarder ce que devient l’ nergie du vide ou la constante cosmologique apr s renormalisation. On parle bien videmment du vide de la th orie supersym trique sous-jacente. Les corrections boucle sur l’ nergie du vide sont quadratiquement divergentes dans une th orie générique de supergravit , ce qui sugg re une contribution à l’énergie du vide de l’ordre de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ et donc $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{O}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{32}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}`$. Cette contribution est supprimée dans certains mod les, qui ont une correction une boucle de l’ordre de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{O}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{64}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$. Pourtant, il faudrait sans doute une sym trie suppl mentaire pour amener la constante cosmologique au niveau requis, soit $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{123}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}`$. Ceci nous conduit maintenant à discuter de la th orie des cordes, qui est notre meilleur candidat pour une Th orie de Tout incluant la gravit . ### 4.3 Vers une "Th orie de Tout" #### 4.3.1 Les probl mes de la gravit quantique Un des l ments fondamentaux qui manquent notre compr hension de l’Univers et des interactions fondamentales est l’unification des deux grandes th ories du XX me si cle : la relativit g n rale et la m canique quantique. Ecrire une th orie unifi e est un des enjeux majeurs de notre si cle. La solution du probl me de la constante cosmologique devra se situer dans le cadre d’une telle "Th orie de Tout". La gravitation chappe la th orie quantique surtout car des infinis incontr lables et non renormalisables apparaissent lorsque l’on veut calculer des diagrammes contenant des boucles avec des gravitons. Ces corrections sont des puissances qui divergent de plus en plus rapidement lorsqu’on augmente l’ordre du calcul perturbatif, parce que la constante de couplage de la gravitation poss de une dimension. Il existe aussi des probl mes non perturbatifs qui mergent lorsque l’on cherche quantifier la gravitation. Ces probl mes sont pour la premi re fois apparus lorsque les physiciens se sont int ress s aux trous noirs. Le trou noir est une solution non perturbative des quations de la Relativit G n rale dans laquelle la courbure de l’espace-temps liée aux forces gravitationnelles devient très importante : aucune particule ne peut sortir de l’horizon qui l’entoure. L’existence de cet horizon est liée aux questions d’entropie et de température des trous noirs. En effet, la masse d’un trou noir est proportionnelle la surface de son horizon, qui se situe au rayon de Schwarzschild: $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}`$ : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (4.81) o $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}`$ est la surface et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}`$ l’entropie. Lorsque la masse d’un trou noir augmente, sa surface augmente, et aussi son entropie. Proche de l’horizon du trou noir, les effets quantiques créent des paires de particules, dont une est virtuelle et se trouve à l’intérieur du trou noir tandis que l’autre est réelle et se trouve à l’extérieur de l’horizon. Cette dernière est rayonnée par le trou noir : c’est que l’on appelle le rayonnement d’Hawking. Cette radiation est stochastique, avec tous les charactéristiques d’une radiation thermique émise par un corps noir. Cet effet est à rapprocher de l’effet Unruh, selon lequel un observateur dans un référentiel accéléré detecte une radiation thermique. Dans le cas d’un trou noir, sa temp rature est liée à sa masse : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pi }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (4.82) L’entropie du trou noir reflète la perte d’information à travers son horizon, et l’état thermique de la radiation de Hawking, d’une nature stochastique, en est l’expression. Pour la décrire, il faut utiliser un état quantique mixte. Or, on peut imaginer la préparation d’un trou noir à partir d’un état pur. Donc il faut envisager une transition entre un état mixte et un état pur, ce qui n’est pas admis par la mécanique quantique habituelle. On peut voir ce problème déjà au niveau de la radiation de Hawking : consid rons un tat quantique pur composé de deux particules $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}`$ dans une superposition d’états individuels : $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}`$. Si la particule $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}`$ tombe dans le trou noir et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}`$ s’ chappe, l’information tenue par la particule $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}`$ est perdue : $$\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\underset{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{>}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{|}$}$$ (4.83) et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}`$ merge dans un tat mixte! Or, la m canique quantique ne permet pas l’ volution d’un tat pur en tat mixte. Cette discussion met en vidence un conflit entre la th orie quantique et la Relativit G n rale. Au moins un de ces deux piliers de la physique du XXème siècle doit tre modifi . Les physiciens des particules pr f rent modifier la Relativit G n rale en l’ levant la th orie des cordes. Nous continuons avec cette th orie. #### 4.3.2 Introduction la th orie des cordes Comme nous avons déjà vu, un des probl mes majeurs de la gravit quantique est la pr sence d’un nombre illimité de divergences. Ces divergences peuvent tre reli es l’absence de cut-off (coupure) courtes distances dans les th ories de champs usuelles o les particules sont ponctuelles. En effet, on peut rapprocher de mani re infinie proche des objets ponctuels, donnant ainsi naissance des interactions infinies: $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{/}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\left(}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{x}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{6}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\right)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (4.84) On peut adoucir voire effacer ces divergences si l’on impose un cut-off naturel. Pour faire ceci, il suffit de consid rer des objets tendus plut t que ponctuels pour r duire voire supprimer ces divergences. L’option la plus simple est de remplacer les particules ponctuelles par des objets unidimensionnels : les cordes. Les lignes d’univers associ es ces objets tendus deviennent des surfaces d’Univers (world-sheets). Si les cordes sont ferm es, on obtient des tubes dans l’espace-temps, qui forment des circuits de plomberie lorsqu’ils interagissent entre eux, comme indique la figure 4.4. On peut imaginer tendre ce principe a des objets avec davantage de dimensions comme des membranes, dont les lignes d’Univers se transforment en "volumes d’Univers". Nous reviendrons ces objets lorsque nous parlerons des aspects non perturbatifs de la th orie. Historiquement les th ories de cordes furent introduites pour d crire les interactions fortes avant que ne soit d velopp e la QCD. Les tats li s l’ taient par l’interm diaire de cordes possèdant une certaine tension. Les amplitudes des processus lors desquels deux particules donnent $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ particules pouvaient tre directement d riv es d’une th orie quantique des cordes. Avec l’av nement de la QCD, la th orie des cordes fut presque oubli e comme th orie des interactions fortes <sup>17</sup><sup>17</sup>17Quelques th oriciens continuent à travailler sur ce sujet.. En plus, il a été reconnu que l’unitarit nécessitait que les cordes ferm es soient pr sentes dans la th orie. Il apparaissait que le spectre des tats quantiques de la corde ferm e incluerait une particule de spin-2 et sans masse, ce qui tait un inconv nient pour une théorie de l’interaction forte. Pourtant, cela fournit l’id e que les cordes pouvait tre une Th orie de Tout: en effet, une particule de spin 2 sans masse peut tre interpr t e comme un graviton, et dans ce cas la tension de la corde deviendrait beaucoup plus élevée : $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{O}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Le spectre d’excitation des cordes fournit un nombre infini de diff rentes particules. Puisque les cordes se propagent sur une surface d’Univers, le formalisme est bidimensionnel. Les vibrations des cordes peuvent tre d crites en terme d’ondes qui propagent dans les deux sens autour de la corde fermée, vers la gauche ou vers la droite : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{L}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\varphi }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (4.85) Si la corde est ferm e, les ondes gauches et droites sont ind pendantes. Apr s quantification, on peut r crire la th orie comme une th orie de champs deux dimensions, car on travaille sur des surfaces d’Univers. Compar e une th orie quatre dimensions, il est relativement simple d’obtenir une th orie finie. Dans ce cas, la th orie poss de une sym trie conforme qui est d crite par un groupe de sym trie de dimension infinie deux dimensions. Cette sym trie classique ne doit pas tre bris e par des anomalies lorsqu’on rend la th orie quantique. Annuler l’anomalie conforme implique que dans un espace-temps plat les ondes gauches et droites sont toutes les deux quivalentes $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{26}$}`$ bosons si la th orie n’a pas de supersym trie et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}`$ supermultiplets de bosons et fermions si la th orie possède une supersym trie N=1 sur la surface d’Univers. #### 4.3.3 Les grandes classes de th ories de cordes Parmi les mod les de th ories de cordes consistantes, on trouve la corde bosonique, une th orie en $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{26}$}`$ dimensions qui ne contient pas de fermions. La corde bosonique a un vide instable, c’est- -dire qu’un espace-temps plat est instable. On a aussi des mod les de supercordes consistants en $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}`$ dimensions, qui contiennent des fermions et poss dent des vides stables, donc des espaces-temps plats stables, mais n’ayant pas de fermions chiraux. Le mod le de corde h t rotique en $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}`$ dimensions est aussi un mod le supersym trique, mais il poss de en plus du mod le pr c dent des fermions chiraux, car la violation de la parit a t int gr e puisque les ondes gauches et droites sont trait es diff remment. Cette th orie est appel e h t rotique car elle est contruite partir d’une th orie de supercordes pour la partie fermionique et d’une th orie de cordes bosoniques $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{26}$}`$ dimensions pour la partie bosonique. Les $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{16}$}`$ dimensions suppl mentaires qui sont autant de degr s de libert suppl mentaires sont per ues comme des champs suppl mentaires dans la th orie $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}`$ dimensions. Enfin, on peut citer les mod les de cordes h t rotiques en $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}`$ dimensions qui sont obtenus en compactifiant les six dimensions supplémentaires des rayons de l’ordre de la longueur de Planck, ou alors en travaillant directement $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}`$ dimensions en rempla ant les dimensions manquantes par d’autres degr s de libert internes comme des fermions ou des groupes de sym trie. De cette mani re, il a t possible d’incorporer un groupe de jauge unifié et m me des mod les ressemblant au Mod le Standard. ##### Les th ories de supercordes Comme nous l’avons dit, le mod le de corde bosonique poss de beaucoup plus de d savantages que les autres mod les: il a $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{26}$}`$ dimensions, pas de fermions et un vide instable! Les physiciens se sont donc concentr s sur des mod les supersym triques donc de supercordes. Or, il existe cinq th ories de supercordes : \- le type IIA qui se r duit basse nergie une supergravit non chirale $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}`$ dimensions ; \- le type IIB qui se r duit basse nergie une supergravit chirale $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}`$ dimensions ; \- la th orie h t rotique $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{8}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\times }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{8}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ qui se r duit basse nergie une supergravit $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}`$ coupl e une th orie de Yang-Mills avec le groupe de jauge $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{8}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\times }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{8}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ ; \- la th orie h t rotique $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{O}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{32}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ qui se r duit basse nergie une supergravit $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}`$ coupl e une th orie de Yang-Mills de groupe de jauge $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{O}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{32}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ ; \- le type I qui contient la fois des cordes ouvertes et ferm es et qui se r duit une supergravit $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}`$ coupl e une th orie de Yang-Mills de groupe de jauge $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{S}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{O}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{32}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Chacune de ces th ories est diff rente de l’autre. La th orie type I est la seule qui contienne la fois des cordes ouvertes et ferm es alors que les autres ne contiennent que des cordes ferm es. De plus, la structure de jauge des cinq th ories est sensiblement diff rente dans leur structure basse nergie. Nous sommes donc en pr sence de cinq th ories qui permettent apparemment de d crire la gravit comme une force quantique et qui sont diff rentes. Comment comprendre cela ? Existe-t-il un lien entre les diff rentes th ories? #### 4.3.4 La structure non-perturbative de la th orie Dans la construction des th ories de cordes, les physiciens ont commenc par travailler sur une th orie de premi re quantification, c’est- -dire une th orie o l’on quantifie la position et l’impulsion. Dans cette th orie, les interactions sont introduites la main et ne peuvent pas d river d’une seule action. L’unitarit doit tre v rifi e pas pas. La formulation est sur couche de masse. La formulation de la th orie est perturbative. La th orie de seconde quantification (ou th orie des champs) des cordes quantifie les champs. Les interactions y sont explicites dans l’action, l’unitarit est en principe garantie si le Hamiltonien est hermitien, la th orie peut tre formellement crite perturbativement ou non perturbativement, et la th orie poss de une formulation hors couche de masse. Le grand probl me de la th orie des champs de cordes est que, bien que sa formulation soit ind pendante d’une th orie de perturbation, elle est actuellement trop difficile pour tre r solue dans une r gion non perturbative (forts couplages). L’id e cl pour sonder et comprendre la structure non perturbative de la th orie est d’utiliser la dualit . La notion de dualit peut tre comprise facilement en consid rant la th orie du champ magn tique et du champ lectrique. Les quations de Maxwell sont invariantes par les transformations de dualit suivantes : l’on change $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}`$ en $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}`$ et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}`$ en $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}`$ et l’on change $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}`$ en $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}`$, o $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}`$ est la charge d’un monopole magn tique. Or, dans le théorie quantique $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pi }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ o $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ est un entier. On conclut donc que la th orie du champ lectrique d fini dans une r gion fort couplage ($`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}`$ grand) est quivalente une th orie du champ magn tique o le couplage est faible ($`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}`$ petit) et vice-versa. On voit sur cet exemple que l’on peut sonder la th orie lectrodynamique fort couplage (donc difficile r soudre) en tudiant une th orie de monopoles magn tiques faible couplage. C’est exactement le m me principe que l’on applique aux th ories de cordes. Les diff rentes th ories de cordes sont reli es par un r seau de relations de dualit qui permet de comprendre la r gion non perturbative d’une th orie en utilisant la zone perturbative d’une autre th orie. D’autre part, ces relations de dualit laisse supposer qu’une th orie m re gouverne toutes les th ories de cordes comme dans le cas de l’ lectromagn tisme: on a la th orie du champ magn tique reli e la th orie du champ lectrique par dualit , cela implique l’existence d’une th orie regroupant les deux th ories : l’ lectromagn tisme. La th orie m re de ces th ories des supercordes est appel e la M-th orie. La lettre $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}`$ pourrait avoir le sens de mère, ou peut signifier le mot membrane, ou bien magique ou mystérieuse, et certains y entendent le mot $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{a}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{r}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}`$, d’après une autre formulation de la th orie. Mais tout d’abord, citons quelques cons quences de l’existence de la dualit . Des objets tendues de dimensions sup rieure un, appel es membranes, apparaissent lorsqu’on tudie des solutions non perturbatives des th ories de supercordes, et aussi de supergravit en dimension $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}`$ ou $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{11}$}`$. Ce sont des solutions classiques des quations du mouvement : $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{o}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{l}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{i}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{t}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{o}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}`$ qui ont des masses: $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (4.86) Il est clair que ces membranes deviennent des tats l gers lorsqu’on augmente le couplage $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}`$. On peut, par dualit , tudier le r gime de couplage fort de ces objets dans une certaine th orie en tudiant un objet dual dans une autre th orie o le couplage est faible et o les calculs sont (en principe!) faisables puisque perturbatifs. Parmi ces solitons , solutions non perturbatives des th ories de supercordes, on trouve une classe particuli re appel e membranes de Dirichlet ou D-membranes: les extr mit s de cordes ouvertes y sont rattach es. Un exemple de l’utilisation de ces membranes est le calcul de l’entropie d’un trou noir dont nous avons discut pr alablement. Le trou noir est d crit par une D-membrane sur laquelle sont agglutin es des cordes. On parle d’ailleurs de boules de cordes. Pour calculer l’entropie d’un trou noir, il suffit donc de calculer le nombre d’ tats de la D-membrane. La dualit simplifie le calcul. L’existence d’un trou noir implique un couplage fort, mais on peut se ramener au calcul des diff rents tats d’une corde couplage faible (ce qui est bien connu) et par dualit revenir aux diff rents tats de la membrane couplage fort. Pour certains types de trous noirs et pour certaines th ories de cordes, on retrouve la formule d’entropie macroscopique de Hawking et Bekenstein. Enfin, le paradoxe dont nous parlions propos de l’ tat pur qui devient mixte est maintenant r gl . En effet, le système de deux particules $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{A}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{B}$}`$ qui tait dans un tat pur et qui semblait voluer vers un tat mixte peut tre maintenant per u comme un tat intriqu avec l’ tat du trou noir. Or, l’ tat quantique de ce dernier n’est pas facilement accessible physiquement puisque l’on ne peut pas sortir d’informations d’un trou noir. ##### La M-th orie Les relations de dualit entre les cinq th ories de supercordes impliquent que ces cinq th ories ne sont pas autre chose que des solutions diff rentes d’une seule th orie, la M-th orie. Dans cette approche, les th ories de supercordes ne sont rien d’autre que des d veloppements autour des diff rents vides (un pour chaque th orie) d’une m me th orie. La M-th orie vit, quant elle, dans un espace-temps $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{11}$}`$ dimensions: ce serait la limite couplage fort de la th orie IIA et de l’h t rotique $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{8}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\times }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{8}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. En compactifiant la M-th orie sur un cercle on retrouve la th orie type IIA , en la compactifiant sur un cercle divis e par $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{Z}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$, on retrouve la th orie h t rotique $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{8}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\times }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{8}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. On ne connait pas l’action de cette M-th orie, mais on sait que la th orie effective aux basses nergies de la M-th orie est la supergravit $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{11}$}`$ avec $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{N}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. La th orie IIA et la M-th orie sont reli es par la S-dualit qui associe une observable $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ de la th orie IIA une observable $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ de la M-th orie de sorte que $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{f}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$. Le r sultat important qui ressort de la S-dualit appliqu e la th orie type IIA est l’existence de la M-th orie en 11 dimensions avec un couplage directement reli la dimension de la onzi me dimension. Ainsi, on obtient : $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{11}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^{\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{3}$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (4.87) La onzi me dimension n’appara t visible que dans la limite de couplage fort. C’est la raison pour laquelle cette dimension suppl mentaire n’avait pas t per ue avant l’av nement de la dualit : en n’ayant acc s qu’au domaine perturbatif, il tait impossible de la d celer. Comme la taille de la onzi me dimension est li e inversement au couplage, l’unification des couplages est fonction de la taille de cette onzi me dimension. Or, il se trouve que pour que les couplages s’unifient, la taille de cette onzi me dimension doit tre sup rieure celles des six autres dimensions suppl mentaires. En compactifiant les six dimensions une chelle de l’ordre de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}}`$ donc $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{16}$}}`$ GeV , on obtient une taille de l’ordre de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{12}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{13}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ pour la onzi me dimension (appel e $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{11}$}}`$) pour que l’unification des quatre couplages ait lieu $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}}`$. Ainsi, des nergies comprises entre $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{12}$}}`$ Gev et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{16}$}}`$ GeV, le monde appara t de dimension cinq. On peut repr senter l’ tape interm diaire cinq dimensions par deux plateaux de dimensions $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}`$ mais qui, haute nergie, ont chacun $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}`$ dimensions: les six autres ont t compactifi es. Les deux plateaux ou membranes sont s par es par la onzi me dimension. Dans une autre approche, on unifie d’abord les trois couplages des groupes de jauge $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}}`$, et puis le couplage gravitationnel rejoint les trois autres couplages une chelle d’ nergie plus haute $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{11}$}}`$. Dans ce cas, on trouve que $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{11}$}}`$ est environ deux fois plus grand que $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}}`$ et que $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{11}$}}`$ est de l’ordre de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{14}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{15}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{e}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{V}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. Un lien peut maintenant tre fait avec la brisure de supersym trie dans un secteur cach : notre monde et tous ses champs sont contraints de rester sur une des deux membranes, et le secteur cach est sur l’autre membrane. Seuls les champs gravitationnels (nous parlons au pluriel car on trouve le graviton, le gravitino et le dilaton) peuvent se propager dans le bulk, c’est- -dire dans la dimension supplémentaire, et la supersym trie est bris e dans le secteur cach , c’est- -dire sur la membrane cach e. La brisure est ensuite transmise au secteur visible par les champs gravitationnels. Ainsi, des m canismes fond s sur des mod les de supergravit $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{d}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{11}$}`$ coupl e une th orie de Yang-Mills avec groupe de jauge $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{8}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\times }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{8}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ sont propos s pour d crire la brisure de supersym trie. Le secteur cach correspond un groupe de jauge $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{8}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ et le secteur visible l’autre. On peut citer le m canisme de condensation de gauginos qui brise la supersym trie dans le secteur cach : lorsqu’on diminue l’ nergie et qu’on passe une certaine valeur, une interaction forte du secteur cach fait condenser deux gauginos et brise la supersym trie. Nous conclurons ce paragraphe par citer ce qu’il reste accomplir. Le plus grand travail sera d’ crire l’action de la M-th orie et quantifier la th orie. On suppose, cause de sa th orie basse nergie - la supergravit $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{11}$}`$ dimensions - que la th orie contient des 2-membranes et des 5-membranes qui apparaissent comme des cordes ferm es plus basses dimensions, mais actuellement on ne sait pas quantifier l’action d’une membrane avec les techniques connues. D’autre part, on ne sait pas clairement comment ces membranes interagissent bien que, gr ce la dualit , on puisse donner les excitations des membranes en termes de cordes. La M-th orie ne peut pas dans l’ tat actuel de son d veloppement pr dire la valeur de la constante cosmologique ou expliquer pourquoi la supersym trie est bris e. Enfin, la M-th orie devra une fois crite permettre de d terminer l’ nergie d’unification sans avoir ajuster la taille de la onzi me dimension: elle devra donc permettre de pr dire la fois la taille de cette dimension mais aussi la valeur de l’ nergie d’unification des couplages. #### 4.3.5 Compactifications, pr dictions et limites des th ories de supercordes Une pr diction importante de la th orie perturbative des cordes est la valeur de l’ nergie d’unification: $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{U}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{O}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\times }$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{P}$}}{\sqrt{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{8}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\pi }$}}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{few}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\times }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{17}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\text{GeV}}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{.}$}$$ (4.88) Cette valeur est environ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{20}$}`$ fois plus grande que la valeur calculée par une approche qui monte depuis les basses nergies vers les hautes nergies … Pourtant, il est impressionnant que l’estimation (4.88) soit si proche d’un calcul qui traverse $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{14}$}`$ ordres de grandeur des basses nergies vers les hautes! Comme nous l’avons d j fait remarquer, dans la M-theorie, la onzi me dimension peut tre r gl e de mani re ce que tous les couplages s’unifient $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{16}$}}`$ GeV. Il est cependant important de constater que deux approches diff rentes, la th orie quantique des champs et sa renormalisation d’une part et la th orie des cordes d’autre part conduisent, des valeurs de l’ nergie d’unification qui sont si proche. Parmi les limitations de la th orie des cordes, on trouve le probl me de la compactification donc, finalement, le lien avec notre monde $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}`$ dimensions. Les moyens de compactifier la th orie, c’est-à-dire, rendre invisibles les dimensions supplémentaires, sont innombrables. On a beaucoup d’options pour retrouver le mod le standard plus d’autres interactions non observ es. Comment choisir parmi toutes les possibilit s pour compactifier? Est-ce qu’il faut choisir une variété, ou bien un espace avec une géométrie généralisée, un orbifold, par exemple? La première option étudiée en détail afin de maintenir la supersymétrie était de compactifier sur une variété de Calabi-Yau. Or, toute compactification sur un espace Calabi-Yau est param tris e par les champs de moduli qui décrivent les possibles d formations de ces espaces : a priori, les champs de moduli n’ont pas de valeur d termin e. Cela revient se poser la question de savoir quel est l’ tat de vide de la th orie $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}`$ dimensions. L’ nergie du vide $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}`$ dimensions devrait correspondre correspondre la constante cosmologique. Or, il y a un nombre norme d’ tats de vide possibles apr s compactification $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}`$ dimensions qui pourraient correspondre cette nergie: le probl me de la d g n rescence du vide reste à resoudre. ### 4.4 Les dimensions suppl mentaires #### 4.4.1 Pourquoi (pas) des dimensions suppl mentaires? Les dimensions suppl mentaires sont pr sentes dans les th ories de cordes comme nous l’avons vu. Pourtant, l’id e d’un espace-temps plus de quatre dimensions date des ann es $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1920}$}`$: Kaluza et Klein furent les premiers introduire une dimension suppl mentaire dans le but d’unifier lectromagn tisme et gravitation. L’ lectromagn tisme pouvait s’expliquait g om triquement comme la gravitation si l’on ajoutait une dimension suppl mentaire (une ligne et une colonne de plus la matrice repr sentant le tenseur m trique). En compactifiant cette dimension, ils montr rent que l’on retrouvait les lois de l’ lecromagn tisme telles qu’elles sont dans notre monde quatre dimensions. Les th ories modernes proviennent plus ou moins de la M-th orie o deux membranes sont s par es par une dimension suppl mentaire. Pourtant, ces th ories s’en loignent plus ou moins fortement. En effet, la M-th orie et sa version basse nergie - la supergravit en dimension $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{11}$}`$ coupl e une th orie de Yang-Mills de groupe de jauge $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{8}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\times }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{E}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{8}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}`$ \- semble n’autoriser que les champs gravitationnels se propager dans toutes les dimensions et contraint les champs qui font notre monde sur une membrane et les champs du secteur cach sur l’autre membrane; de plus, la dimension de la onzi me dimension est d termin e par l’unification des couplages. Alors que dans les th ories de Kaluza et Klein, on peut trouver diff rentes tailles pour la (ou les) dimension(s) suppl mentaire(s), et toutes les variantes possibles pour la localisation des champs, y comprise la propagation des champs de jauge dans des dimensions suppl mentaires. #### 4.4.2 Les diff rents types de mod les Il existe de nombreux mod les diff rents. Les deux classes principales sont celle o la dimension suppl mentaire est plate et celle o elle est courbe. Dans le premier cas, on trouve des mod les avec dimension suppl mentaire plate grande, c’est- -dire pouvant aller jusqu’ $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0.1}$}`$ mm qui est la limite haute de taille de dimension suppl mentaire d’apr s des exp riences de type Cavendish. La taille de la ou des dimensions suppl mentaires affecte directement la masse de Planck $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}`$ dans l’espace $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{4}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}`$ dimensions: $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{p}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (4.89) o R est la taille de la dimension suppl mentaire. Par exemple, dans le cas o $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}`$ et si l’on se fonde sur une th orie de cordes pour crire que $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{y}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{n}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (4.90) avec $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{M}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{s}$}`$ de l’ordre du TeV, on obtient des limites sur la taille des dimensions suppl mentaires par la cosmologie qui donnerait $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{10}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{8}$}`$cm si seule la gravit se propage dans les dimensions suppl mentaires. Dans chacun des mod les, on peut choisir d’autoriser certains champs se propager dans toutes les dimensions et d’autres rester localisés sur les membranes. Parmi les mod les avec une dimension suppl mentaire courbe, citons le mod le de Randall-Sundrum, où il y a deux membranes. Les champs du Mod le Standard sont localis s sur une membrane et la gravitation se propage partout. Randall et Sundrum ont montr que l’on pouvait red finir le probl me de hi rarchie entre la masse de Planck et l’ chelle lectrofaible sans avoir faire appel la supersym trie. Il existe cependant des mod les supersym triques de ce type o la supersym trie est bris e par des anomalies. #### 4.4.3 Signatures exp rimentales Il est possible de d celer la pr sence de dimensions suppl mentaires l’ chelle du TeV au LHC. Pour cela, il faut bien videmment que les particules en jeux puissent sentir la pr sence de la (ou des)dimension(s) suppl mentaire(s). Les signatures typiques sont des excitations dites de Kaluza-Klein ; des r sonnances dans l’ volution des sections efficaces des r actions en fonction de l’ nergie peuvent apparaitre. Elles sont dues au "tours" d’excitations de Kaluza-Klein. On peut comprendre ces "tours" en faisant une analogie avec une particule qui rencontre un puit de potentiel en m canique quantique: son nergie va tre quantifi e dans le puit fournissant une tour d’excitations. Dans notre cas, la dimension suppl mentaire joue le r le du puit. D’autres signatures pourraient tre fournies par la cr ation d’un micro trou noir. En effet, tout objet peut se transformer en trou noir s’il est contract en de a de son rayon de Schwarschild qui est d fini par $`\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{m}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}^\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}`$. Bien videmment, plus la masse est grande, plus l’on peut obtenir facilement le rayon de Schwarschild. Or, comme on l’a vu, les dimensions suppl mentaires peuvent augmenter la valeur de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{G}$}`$. Donc s’il y a des dimensions suppl mentaires de taille suffisante, on peut esp rer avec les nergies en jeu au LHC passer en de a du rayon de Schwarschild et cr er des micro trous noirs. Ceux-ci devraient s’ vaporer rapidemment, puisque le rayonnement de Hawking implique que le trou noir rayonne de mani re inversement proportionnelle sa masse. Des études par les collaborations ATLAS et CMS ont démontré que ce rayonnement de Hawking serait sans doute visible au LHC. #### 4.4.4 Des dimensions en moins haute nergie? Tandis que ce n’est pas prévu dans le théorie des cordes, il est possible qu’au moins quelques-unes des dimensions de l’espace-temps paraissent discrètes hautes nergies. Le principe est d’imposer un cut-off non par la dimension donn e aux particules comme c’est le cas en th orie des cordes, mais par la structure interne de l’espace-temps. Comme dans les théories de champs sur le r seau, le mouvement est d compos en une succession de bonds d’un point discret de l’espace un autre. Aux basses énergies, la résolution expérimentale n’est pas suffisante pour détecter ces bonds, et une dimension discrète ressemble à une dimension continue. Pourtant, aux hautes énergies, cette dimension disparaît, et la dimensionalité de l’espace-temps diminue! ### 4.5 L’ nergie noire Nous terminons ce cours avec un des plus grands myst res de la Physique contemporaine: la myst rieuse nergie noire. Commen ons par nous rappeler des quations de la Relativit G n rale: $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (4.91) $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}`$ est le tenseur de Ricci (ou de courbure) , $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}`$ sa trace et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}`$ est le tenseur d’ nergie-impulsion. Ces quations relient la g om trie de l’espace-temps la pr sence de mati re (ou d’ nergie). Pourtant, cette forme n’est pas la forme la plus g n rale. En effet, cette quation vient de l’int gration d’une quation. On a la conservation du tenseur nergie-impulsion: $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$ o $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}`$ est la d riv e covariante et on a aussi $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{(}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{)}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$. En int grant cette derni re quation, on constate qu’il est possible de rajouter un terme proportionnel au tenseur m trique $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}`$. En effet, on a toujours $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}^{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0}$}`$. La forme la plus g n rale des quations de la Relativit G n rale est donc: $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{2}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{R}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{+}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{k}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{T}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{,}$}$$ (4.92) o $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}`$ est la constante cosmologique. Les mesures actuelles des Supernovae lointaines tendent prouver que l’énergie dans le vide n’est pas nulle, et pourrait correspondre à une constante cosmologique. En effet, ces mesures et d’autres montrent que l’Univers est dans une phase d’expansion qui est acc l r e. Or, si l’on reprend les quations de la Relativit G n rale, et qu’on place la partie contenant $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}`$ du c t du tenseur nergie-impulsion et qu’on interpr te le terme $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{g}$}_{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mu }$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\nu }$}}`$ comme un terme d’ nergie-impulsion homog ne et isotrope, on trouve qu’il correspond un fluide tr s particulier qui a une densit d’ nergie positive et une pression n gative. Ce fluide qui remplit tout l’Univers de mani re homog ne permet d’expliquer la phase d’acc l ration de l’Univers. La pression n gative permet l’expansion acc l ree de l’Univers. Ce fluide a re u, cause de sa propri t inconnue jusqu’alors de pression n gative, le nom de quintessence ou cinqui me l ment. C’est aussi ce que l’on nomme l’ nergie noire. Si $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Lambda }}$}`$ est une vraie constante, le rapport pression sur densit du fluide de quintessence devrait tre gale $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}`$. Il se trouve que les mesures actuelles donnent un rapport pression sur densit de $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{<}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{0.8}$}`$. Les mesures r alis es d’apr s le CMB montre que l’Univers est plat c’est- -dire que $$\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\mathrm{\Omega }}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\frac{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}}{\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{=}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{1}$}$$ o $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}`$ est la densit totale d’ nergie-mati re de l’Univers et $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\rho }$}_\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{c}$}`$ est la densit critique qui correspond un Univers plat. Les mesures sur les Supernovae combin es au fait que l’on sait que l’Univers est plat indique que $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{70}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\%}$}`$ de la densit d’ nergie-mati re de l’Univers est de l’ nergie noire. Les $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{30}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\%}$}`$ restant sont les $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{25}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\%}$}`$ de mati re noire non baryonique et les $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\%}$}`$ de mati re baryonique (il est noter que sur ces $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{5}$}\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{\%}$}`$, la mati re visible sous forme d’ toiles ne repr sente qu’un dixi me!). Qu’est-ce que l’ nergie noire? Si on la d crit par un champs scalaire, quel particule correspond-il? Cela reste myst rieux. De plus, comme nous l’avons d j dit dans le chapitre sur la brisure de supersym trie, si cette nergie correspond l’ nergie du vide, comment obtenir le bon vide? A l’heure actuelle, aucune th orie ne peut pr dire la valeur de la constante<sup>18</sup><sup>18</sup>18On utilise le mot constante par habitude mais comme on l’a vu il se pourrait que cette constante n’en soit pas une. cosmologique, donc aucune th orie ne poss de le v ritable tat de vide. C’est le grand d fi contemporain. La "Th orie du Tout" ne pourra tre que la th orie qui pr dit la valeur de l’ nergie noire. Bibliographie g n rale Mod le Standard et brisure lectrofaible $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ Le cours cette cole de Gif 2004 de D.Treille : Le Mod le Standard. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ S. Weinberg, “The quantum theory of fields. Vol. 2: Modern applications”, (Cambridge Univ. Press, Cambridge, UK) (1996). $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ J.M.Fr re, “La th orie lectrofaible et au-del ”, 22 me cole Joliot-Curie de Physique nucl aire 2003, "Interaction faible et noyau : l’histoire continue….", p27. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ C. T. Hill and E. H. Simmons, “Strong dynamics and electroweak symmetry breaking”, Phys. Rept. 381 (2003) 235 \[Erratum-ibid. 390 (2004) 553\] \[arXiv:hep-ph/0203079\]. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ E. Witten, “When symmetry breaks down”, Nature 429 (2004) 507. Supersym trie : $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ S. P. Martin, “A supersymmetry primer”, arXiv:hep-ph/9709356. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ G. L. Kane, “Weak scale supersymmetry: A top-motivated-bottom-up approach”, arXiv:hep-ph/0202185. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ M. F. Sohnius, “Introducing Supersymmetry,” Phys. Rept. 128 (1985) 39. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ A.Deandrea, “Interactions lectrofaibles et introduction la supersym trie”, disponible l’adresse : http://deandrea.home.cern.ch/deandrea/seminars/ew.ps $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ J. R. Ellis,“Supersymmetry for alp hikers”, . Physique des neutrinos : $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ W. Buchmuller, “Neutrinos, grand unification and leptogenesis”, arXiv:hep-ph/0204288. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ J. R. Ellis, “Particle physics and cosmology”, arXiv:astro-ph/0305038. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ J.R.Ellis, “Limits of the standard model”, . $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ S. F. King, “Neutrino mass models”, Rept. Prog. Phys. 67 (2004) 107 \[arXiv:hep-ph/0310204\]. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ M. C. Gonzalez-Garcia and Y. Nir, “Developments in neutrino physics”, Rev. Mod. Phys. 75 (2003) 345 \[arXiv:hep-ph/0202058\]. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ D.Vignaud, “Comment la masse vint aux neutrinos”, 22 me cole Joliot-Curie de Physique nucl aire 2003, Interaction faible et noyau : l’histoire continue…., p289. Grande Unification, supergravit , th ories des cordes $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ T. P. Cheng and L. F. Li, “Gauge Theory Of Elementary Particle Physics”, (Oxford Science Publications, Clarendon Press, Oxford, UK) (1984). $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ G. G. Ross, “Grand Unified Theories”, (Frontiers In Physics, Benjamin/Cummings, Reading, USA) (1984). $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ R.N.Mohapatra, “Unification and supersymmetry”, . $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ J. R. Ellis, “Beyond the standard model for hillwalkers”, $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ J. R. Ellis, “Supersymmetry for Alp hikers”, . $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ M. Kaku, “Introduction to superstrings and M-theory”, (Springer, New York, USA) (1999). $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ J. Wess and J. Bagger, “Supersymmetry and supergravity,” (Princeton Univ. Press, Princeton, USA) (1992). $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ S. Weinberg, “The quantum theory of fields. Vol. 3: Supersymmetry,” (Cambridge Univ. Press, Cambridge, UK) (2000) 419. $`\colorbox[rgb]{1,1,1}{$\textcolor[rgb]{0,0,0}{}$}`$ D. Bailin and A. Love, “Supersymmetric gauge field theory and string theory,” (Graduate student series in physics, IOP, Bristol, UK) (1994).
warning/0506/cs0506086.html
ar5iv
text
# Large System Decentralized Detection Performance Under Communication Constraints ## I Introduction This paper considers decentralized detection in energy-constrained, large wireless sensor networks in noisy, band-limited channels. Although there is a considerable amount of previous work on the subject of distributed detection, most of it used to ignore the effect of noisy channels between the local sensors and data fusion center. Even less is the attention received by bandlimited noisy channels in the context of decentralized detection. For example, while distributed detection performance of an energy-constrained wireless sensor network over a noisy channel has been considered recently , it assumes orthogonal sensor-to-fusion center communication leading to an infinite bandwidth assumption. However, in applications involving dense, low-power, distributed wireless sensor networks it is more likely that all nodes will share a common available bandwidth. In this case, the assumption of large sensor systems implies non-orthogonal communication between the sensor nodes and the fusion sensor. An important design objective in low-power wireless sensor systems is to extend the whole network lifetime. Thus, a sensible constraint on the sensor system is a finite total power . In this paper, the bandwidth constraint is taken into account by assuming non-orthogonal direct-sequence code-division multiple-access (DS-CDMA) communication between sensors and the data fusion center. The main contribution of this paper is the derivation of the decentralized detection performance, in closed-form, under a total power constraint when the communication channel between the local sensors and the fusion center is both bandlimited and noisy. As we will see, the performance is a function of the exact signalling codes used by the distributed sensors for any finite-size sensor network. However, in the case of random spreading we are able to derive an elegant and simple closed-form expression that is independent of the exact spreading codes once we consider asymptotically large sensor systems. This is our main result and, as we will see, it allows us to draw general conclusions regarding the design of wireless sensor systems under such total power constraints in noisy and bandlimited channels. The remainder of the paper is organized as follows: In Section II we present our system model. Next, in Section III we use random matrix theory to derive a closed-form expression for the decentralized detection performance in a large sensor system followed by a discussion of our analysis. Finally, in Section IV we conclude by summarizing our results. ## II System Model Description We consider a binary hypothesis testing problem in an $`N_s`$-node wireless sensor network connected to a data fusion center via distributed parallel architecture. Let us denote by $`H_0`$ and $`H_1`$ the null and alternative hypotheses, respectively, having corresponding prior probabilities $`P(H_0)=p_0`$ and $`P(H_1)=p_1`$. We will consider that the observed stochastic process under each hypothesis consists of one of two possible Gaussian signals, denoted by $`X_{0,n}`$ and $`X_{1,n}`$, corrupted by additive white Gaussian noise. Under the two hypotheses the $`n`$-th local sensor observation $`z_n`$, for $`n=1,\mathrm{}N_s`$, can be written as $`H_0:z_n`$ $`=`$ $`X_{0,n}+v_n`$ $`H_1:z_n`$ $`=`$ $`X_{1,n}+v_n`$ (1) where the observation noise $`v_n`$ is assumed to be zero-mean Gaussian with the collection of noise samples having a covariance matrix $`\mathrm{\Sigma }_v`$. Each local sensor processes its observation $`z_n`$ independently to generate a local decision $`u_n(z_n)`$ which are sent to the fusion center. Let us denote by $`𝐫(u_1(z_1),u_2(z_2),\mathrm{},u_{N_s}(z_{N_s}))`$ the received signal at the fusion center. The fusion center makes a final decision based on the decision rule $`u_0(𝐫)`$. The problem at hand is to choose $`u_0(𝐫),u_1(z_1),u_2(z_2),\mathrm{},u_{N_s}(z_{N_s})`$ so that a chosen performance metric is optimized. The solution to this problem is known to be too complicated under the most general conditions . While optimal local processing schemes have been derived under certain special assumptions, a class of especially important local processers are those that simply amplify the observations before retransmission to the fusion center . Thus, the local sensor decisions sent to the fusion center are given by, $`u_n=gz_n`$ for $`n=1,\mathrm{}N_s`$ where $`g>0`$ is the analog relay amplifier gain at each node. In our model all sensor nodes share a common bandwidth and a total available energy. For analytical reasons, as well as due to their practical relation to DS-CDMA communications, we consider bandwidth sharing non-orthogonal communication based on spreading in which each sensor node is assigned a signature code of length $`N`$. If the $`n`$-th sensor node is assigned the code $`𝐬_n`$, the received chip-matched filtered and sampled discrete-time signal at the fusion center can be written as $`𝐫=g_{n=1}^{N_s}𝐬_𝐧z_n+𝐰=g\mathrm{𝐒𝐳}+𝐰`$ where $`𝐫`$ and $`𝐰`$ are $`N`$-dimensional received signal and receiver noise vectors, respectively and the $`n`$-th column of the $`N\times N_s`$ matrix $`𝐒`$ is equal to the vector $`𝐬_n`$. We assume that the receiver noise is a white Gaussian noise process so that the filtered noise vector $`𝐰𝒩(\mathrm{𝟎},\sigma _w^2𝐈_N)`$.Then we have that $`H_0:𝐫`$ $``$ $`𝒩(𝐦_0,\mathrm{\Sigma }_0)`$ $`H_1:𝐫`$ $``$ $`𝒩(𝐦_1,\mathrm{\Sigma }_1)`$ (2) where, for $`j=0,1`$, $`𝐦_j=g𝐒𝔼\{𝐗_j\}`$ and $`\mathrm{\Sigma }_j=g^2𝐒\left(Cov(𝐗_j)+\mathrm{\Sigma }_v\right)𝐒^T+\sigma _w^2𝐈_N`$. To be specific, consider the detection of a deterministic signal so that $`𝐗_1=𝐗_0=m\mathrm{𝟏}`$ is known ($`m>0`$) and $`\mathrm{\Sigma }_0=\mathrm{\Sigma }_1=\mathrm{\Sigma }`$ where ($`\mathrm{𝟏}`$ is the vector of all ones) $`\mathrm{\Sigma }=g^2𝐒\mathrm{\Sigma }_v𝐒^T+\sigma _w^2𝐈_N`$. With these assumptions we also have that $`𝐦_1=𝐦_0=gm\mathrm{𝐒𝟏}`$ and the radiated power of node $`n`$ is then given by $`𝔼\{|u_n|^2\}=g^2𝔼\{|z_n|^2\}=g^2(m^2+\sigma _v^2)`$ where $`\sigma _v^2`$ is the observation noise variance. Let us define the total power constraint the whole sensor system is subjected to as $`P`$, so that the amplifier gain $`g`$ is given by $`g`$ $`=`$ $`\sqrt{{\displaystyle \frac{P}{N_s(m^2+\sigma _v^2)}}}.`$ (3) Then, it can be shown that the optimal threshold rule at the fusion center is of the form $`u_0(\text{r})`$ $`=`$ $`\{\begin{array}{ccccc}1& & & & \\ & \text{if}& T\left(\text{r}\right)& & \tau ^{}\\ 0& & & <& \end{array},`$ (7) where we have defined the decision variable $`T`$ as $`T\left(\text{r}\right)=\left(𝐦_1𝐦_0\right)^T\mathrm{\Sigma }^1\text{r}=2gm\mathrm{𝟏}^T𝐒^T\left(g^2𝐒\mathrm{\Sigma }_v𝐒^T+\sigma _w^2𝐈_N\right)^1\text{r}`$ and $`\tau ^{}`$ is the threshold that depends on the specific optimality criteria. It can be shown that the false-alarm $`P_f`$ and miss $`P_m`$ probabilities of the detector (7) are given by $`P_f`$ $`=`$ $`Q\left({\displaystyle \frac{\tau ^{}+2g^2m^2\mathrm{𝟏}^T𝐒^T\mathrm{\Sigma }^1\mathrm{𝐒𝟏}}{2gm\sqrt{\mathrm{𝟏}^T𝐒^T\mathrm{\Sigma }^1\mathrm{𝐒𝟏}}}}\right),`$ (8) and $`P_m`$ $`=`$ $`Q\left({\displaystyle \frac{2g^2m^2\mathrm{𝟏}^T𝐒^T\mathrm{\Sigma }^1\mathrm{𝐒𝟏}\tau ^{}}{2gm\sqrt{\mathrm{𝟏}^T𝐒^T\mathrm{\Sigma }^1\mathrm{𝐒𝟏}}}}\right).`$ (9) For example, in the case of Neyman-Pearson optimality at the fusion center, $`\tau ^{}`$ is chosen to minimize $`P_m`$ subject to an upper bound on $`P_f`$. On the other hand under Bayesian minimum probability of error optimality one would choose $`\tau ^{}`$ to minimize $`P_e=p_0P_f+p_1P_m`$. As one would expect, the performance of course depends on the particular codes assigned to each sensor node as seen from (8) and (9). Thus, while it is possible to evaluate the performance for specific systems via (8) and (9), it is rather difficult to draw general conclusions regarding the design of decentralized detection systems. However, such conclusions can be reached for large systems through asymptotic analysis, as we show next. ## III Large Sensor System Performance Analysis Let us assume that the spreading codes are chosen randomly so that each element of $`𝐬_n`$ takes either $`\frac{1}{\sqrt{N}}`$ or $`\frac{1}{\sqrt{N}}`$ with equal probability. Moreover, we take independent sensor observations such that $`\mathrm{\Sigma }_v=\sigma _v^2𝐈`$. Let us assume a large sensor system such that both $`N_s`$ and $`N`$ are large such that $`lim_N\mathrm{}\frac{N_s}{N}=\alpha `$. Now using a theorem on the convergence of the empirical distribution of eigenvalues of a large random matrix proven in , we may prove the following proposition, which is the main result of this paper: ###### Proposition 1 With $`𝐒`$ and $`\mathrm{\Sigma }`$ defined as above, $`g^2\mathrm{𝟏}^T𝐒^T\mathrm{\Sigma }^1\mathrm{𝐒𝟏}`$ $``$ $`\left({\displaystyle \frac{\sigma _v^2}{N_s}}+{\displaystyle \frac{m^2+\sigma _v^2}{P\beta _0}}\right)^1,`$ (10) almost surely, as $`N\mathrm{}`$, where $`\beta _0={\displaystyle \frac{\sqrt{\left(\gamma +\sigma _w^2\right)^2\alpha ^2+2\gamma \left(\sigma _w^2\gamma \right)\alpha +\gamma ^2}\left(\gamma +\sigma _w^2\right)\alpha +\gamma }{2\gamma \sigma _w^2}}`$ (11) with $`\gamma =\frac{P}{N}\left(1+\frac{m^2}{\sigma _v^2}\right)1`$ and $`\mathrm{\Sigma }_v=\sigma _v^2𝐈`$. ###### Proof: See Appendix I. The proposition 1 leads to the following corollary on the asymptotically large sensor system performance of decentralized detection in noisy bandlimited channels: ###### Corollary 1 With all notation as defined above, when $`lim_N\mathrm{}\frac{N_s}{N}=\alpha `$, the large sensor network performance of the decentralized detection is given by $`P_fQ\left(\sqrt{\mu }(\tau ^{}+\frac{2m^2}{\mu })/2m\right)`$ and $`P_mQ\left(\sqrt{\mu }(\frac{2m^2}{\mu }\tau ^{})/2m\right)`$ where $`\mu =\frac{\sigma _v^2}{N_s}+\frac{m^2+\sigma _v^2}{P\beta _0}`$. The above corollary leads to insights on large sensor system performance of decentralized detection in noisy, bandlimited channels. For instance, in the special case of minimum probability of error optimality at the fusion center, according to corollary 1, the large system probability of error is asymptotically given by $`P_e(\alpha )`$ $``$ $`Q\left(m/\sqrt{\mu }\right),`$ (12) where convergence is almost surely and $`\mu `$ is as defined above. Figure 1a shows the convergence of the random-spreading based decentralized detection performance as predicted by (12). Note that the exact analysis in Fig. 1a was obtained for a random choice of the code matrix $`𝐒`$. As can be seen from Fig. 1a, (12) provides a good approximation to the detection performance for large spreading lengths $`N`$, and thus for large-sensor systems (since $`N_s=N\alpha `$). More importantly, we can observe from Fig. 1a that for each fixed $`N`$, increasing $`\alpha `$ improves the decentralized detection performance. Since this is equivalent to increasing the number of sensors $`N_s`$ allowed in the system for a fixed bandwidth we conclude that it is better to allow as many sensors to send their local decisions to the fusion center. In fact, for large alpha, one can show that $`\beta _0\frac{1}{\sigma _w^2}`$, and as a result, in this case the error probability in (12) goes to $`P_e(\alpha )`$ $``$ $`Q\left(\sqrt{{\displaystyle \frac{P}{\sigma _w^2}}\left(1+{\displaystyle \frac{\sigma _v^2}{m^2}}\right)^1}\right).`$ (13) On the other hand, if one were to allocate all available power $`P`$ and the total bandwidth to just one sensor node the fusion center performance will be given by $`P_{e,1}`$ $`=`$ $`Q\left(\sqrt{{\displaystyle \frac{P}{\sigma _w^2}}\left({\displaystyle \frac{P/\sigma _w^2}{m^2/\sigma _v^2}}+\left(1+{\displaystyle \frac{\sigma _v^2}{m^2}}\right)\right)^1}\right).`$ (14) Comparison of (13) and (14) shows that allowing more sensor nodes in the network is better even if the channel is both noisy and bandlimited. This comparison is shown in Fig. 1b. First, we can observe from Fig. 1b that as $`N`$ increases the fusion center performance improves. Secondly we see that as $`N\mathrm{}`$, the performance for large $`\alpha `$ indeed goes to (13). Third, Fig. 1b confirms that combining more local decisions is better than allocating all available power and bandwidth to one sensor. Moreover, the performance improves monotonically with increasing $`\alpha `$ (for a fixed $`N`$) showing that it is better to combine as many local decisions as possible at the fusion center. We should divide the available power among all nodes and allow all of them to share the available bandwidth even if they are to interfere with each other due to non-orthogonality. ## IV Conclusions We analyzed the decentralized detection performance of a total average power constrained wireless sensor network in a noisy and bandlimited channel. Assuming that the sensors-to-fusion center communication is based on DS-CDMA, a closed form expression for the fusion performance, and its large system asymptotic under random spreading were derived. It was shown that in a noisy, bandlimited channel it is beneficial to combine as many sensor local decisions as possible even if this leads to non-orthogonal sensor-to-fusion center communication. ## Appendix A The Proof of Proposition 1 ###### Proof: Using the definitions of $`𝐒`$ and $`\mathrm{𝟏}`$, we can write $`g^2\mathrm{𝟏}^T𝐒^T\mathrm{\Sigma }^1\mathrm{𝐒𝟏}=g^2\left({\displaystyle \underset{n=1}{\overset{N_s}{}}}𝐬_n^T\mathrm{\Sigma }^1𝐬_n+{\displaystyle \underset{n=1}{\overset{N_s}{}}}{\displaystyle \underset{\underset{n^{}n}{n^{}=1}}{\overset{N_s}{}}}𝐬_n^T\mathrm{\Sigma }^1𝐬_n^{}\right)`$ (15) Let $``$ denote a set of sensor indices (i.e. $`\{1,2,\mathrm{},N_s\}`$), $`𝐒_𝒜`$ denote the matrix $`𝐒`$ with column indices specified by set $`𝒜`$ deleted, $`𝚲_n=g^2\sigma _v^2𝐈_n`$ and $`𝐐_𝒜=\left(𝐒_𝒜𝚲_{N_s|𝒜|}𝐒_𝒜+\sigma _w^2𝐈_N\right)`$ where $`𝐈_n`$ and $`|𝒜|`$ are the $`n\times n`$ identity matrix and the cardinality of set $`𝒜`$, respectively. Then, for $`n=1,\mathrm{},N_s`$, using the matrix inversion lemma we can show that $`𝐬_n^T\mathrm{\Sigma }^1𝐬_n=𝐬_n^T𝐐_{\{n\}}^1𝐬_n/(1+g^2\sigma _v^2𝐬_n^T𝐐_{\{n\}}^1𝐬_n)`$. But, applying Theorem 7 of and using (3), we can show that $`𝐬_n^T𝐐_{\{n\}}^1𝐬_n\beta _0`$ almost surely, where $`\beta _0`$ is as given by (11) and $`\gamma =\frac{P}{N}\left(1+\frac{m^2}{\sigma _v^2}\right)^1`$. Combining these we have almost surely $`𝐬_n^T\mathrm{\Sigma }^1𝐬_n`$ $``$ $`\left(\beta _{0}^{}{}_{}{}^{1}+g^2\sigma _v^2\right)^1.`$ (16) Similarly, repeated application of matrix inversion lemma twice show that, $`𝐬_n^T\mathrm{\Sigma }^1𝐬_n^{}={\displaystyle \frac{𝐬_n^T𝐐_{\{n,n^{}\}}^1𝐬_n^{}}{\left(1+g^2\sigma _v^2𝐬_n^T𝐐_{\left\{n\right\}}^1𝐬_n\right)\left(1+g^2\sigma _v^2𝐬_n^{}^T𝐐_{\{n,n^{}\}}^1𝐬_n^{}\right)}}.`$ (17) Now the use of Theorem 7 of shows that RHS goes to zero almost surely, for $`nn^{}`$. Substituting (16) and (17) in (15) gives (10), completing the proof.
warning/0506/hep-th0506109.html
ar5iv
text
# References hep-th/0506109 ICTP/June/2005 Nilpotent symmetries for a spinning relativistic particle in augmented superfield formalism R.P.Malik <sup>*</sup><sup>*</sup>* Permanent Postal Address: S. N. Bose National Centre for Basic Sciences, Block-JD, Sector-III, Salt Lake, Kolkata-700 098, India. Electronic Mail Address: malik@boson.bose.res.in. The Abdus Salam International Centre for Theoretical Physics, Strada Costiera 11, 34014 Trieste, Italy Abstract: The local, covariant, continuous, anticommuting and nilpotent Becchi-Rouet-Stora-Tyutin (BRST) and anti-BRST symmetry transformations for all the fields of a $`(0+1)`$-dimensional spinning relativistic particle are obtained in the framework of augmented superfield approach to BRST formalism. The trajectory of this super-particle, parametrized by a monotonically increasing evolution parameter $`\tau `$, is embedded in a $`D`$-dimensional flat Minkowski spacetime manifold. This physically useful one-dimensional system is considered on a three $`(1+2)`$-dimensional supermanifold which is parametrized by an even element ($`\tau `$) and a couple of odd elements ($`\theta `$ and $`\overline{\theta }`$) of the Grassmann algebra. Two anticommuting sets of (anti-)BRST symmetry transformations, corresponding to the underlying (super)gauge symmetries for the above system, are derived in the framework of augmented superfield formulation where (i) the horizontality condition, and (ii) the invariance of conserved quantities on the supermanifold, play decisive roles. Geometrical interpretations for the above nilpotent symmetries (and their generators) are provided. Keywords: Augmented superfield formalism; free spinning relativistic particle; (super)gauge symmetries; reparametrization invariance; nilpotent (anti-)BRST symmetries PACS numbers: 11.15.-q; 12.20.-m; 11.30.Ph; 02.20.+b 1 Introduction For the covariant canonical quantization of gauge theories These theories are endowed with the first-class constraints in the language of Dirac’s prescription for the classification of constraints. The local (non-)Abelian 1-form interacting gauge theories provide an almost exact theoretical basis for the three (out of four) fundamental interactions of nature., one of the most elegant and intuitive approaches is the Becchi-Rouet-Stora-Tyutin (BRST) formalism . In this formalism, the unitarity and “quantum” gauge (i.e. BRST) invariance are very naturally respected at any arbitrary order of perturbative computations for any arbitrary physical process allowed by the interacting gauge theories (where there exists self-interaction as well as the coupling between the (non-)Abelian gauge field and the matter fields). In fact, the whole strength of BRST formalism appears in its full blaze of glory in the context of an interacting non-Abelian gauge theory where the (anti-)ghost fields are required in the precise proof of unitarity. To be more accurate, for every gluon loop (Feynman) diagram, one requires a ghost loop diagram so that unitarity of the theory could be maintained at any given order of perturbative calculation (see, e.g., for details). In modern context, the BRST formalism is indispensable in the realm of topological field theories \[4-6\], topological string theories , string field theories , etc. There are well-known connections of this formalism with the mathematics of differential geometry and supersymmetries. In our present endeavour, we shall be concentrating on the geometrical aspects of the relationship between the BRST formalism and the superfield formalism. To be more elaborate on this topic, it should be noted that, in the framework of usual superfield formulation \[9-14\] of the BRST approach to $`D`$-dimensional $`p`$-form (with $`p=1,2,\mathrm{}..`$) Abelian gauge theories, the gauge theory is considered first on a $`(D+2)`$-dimensional supermanifold parametrized by $`D`$-number of even (commuting) spacetime $`x_\mu `$ variables (with $`\mu =0,1,2\mathrm{}D1)`$ and a couple of odd (anticommuting) Grassmannian variables $`\theta `$ and $`\overline{\theta }`$ (with $`\theta ^2=\overline{\theta }^2=0,\theta \overline{\theta }+\overline{\theta }\theta =0`$). Then, the $`(p+1)`$-form super curvature $`\stackrel{~}{F}^{(p+1)}=\stackrel{~}{d}\stackrel{~}{A}^{(p)}`$ is constructed from the super exterior derivative $`\stackrel{~}{d}=dx^\mu _\mu +d\theta _\theta +d\overline{\theta }_{\overline{\theta }}`$ (with $`\stackrel{~}{d}^2=0`$) and the super $`p`$-form connection $`\stackrel{~}{A}^{(p)}`$ defined on the $`(D+2)`$-dimensional supermanifold. This is subsequently equated, due to the so-called horizontality condition \[9-14\], with the ordinary curvature $`(p+1)`$-form $`F^{(p+1)}=dA^{(p)}`$ defined on the $`D`$-dimensional ordinary spacetime manifold with the exterior derivative $`d=dx^\mu _\mu `$ (with $`d^2=0`$) and the ordinary $`p`$-form connection $`A^{(p)}`$. The above horizontality condition For the 1-form non-Abelian gauge theory, the horizontality condition $`\stackrel{~}{F}^{(2)}=F^{(2)}`$, where 2-form super curvature $`\stackrel{~}{F}^{(2)}=\stackrel{~}{d}\stackrel{~}{A}^{(1)}+\stackrel{~}{A}^{(1)}\stackrel{~}{A}^{(1)}`$ and 2-form ordinary curvature $`F^{(2)}=dA^{(1)}+A^{(1)}A^{(1)}`$, leads to the exact derivation of the nilpotent and anticommuting (anti-)BRST symmetry transformations for the non-Abelian gauge field and the corresponding (anti-)ghost fields of the theory (see, e.g. for details). is christened as the soul-flatness condition in which amounts to setting equal to zero all the Grassmannian components of the (anti-)symmetric super curvature tensor that defines the $`(p+1)`$-form super curvature. We shall be following, in our entire text, the notation adopted in to denote the above general supermanifold as $`(D+2)`$. It should be noted, however, that the modern notation is $`(D,2)`$ for the same. The process of reduction of the $`(p+1)`$-form super curvature to the ordinary $`(p+1)`$-form curvature due to the horizontality condition (i) generates the nilpotent and anticommuting (anti-)BRST transformations for only the gauge fields and the (anti-)ghost fields, (ii) provides the geometrical interpretation for the nilpotent (anti-)BRST charges as the translational generators $`(\text{Lim}_{\overline{\theta }0}(/\theta ))\text{Lim}_{\theta 0}(/\overline{\theta })`$ along the $`(\theta )\overline{\theta }`$-directions of the $`(D+2)`$-dimensional supermanifold, (iii) leads to the geometrical interpretation for the nilpotency property as the two successive translations (i.e. $`(/\theta )^2=(/\overline{\theta })^2=0`$) along either of the Grassmannian directions, and (iv) captures the anticommutativity of the nilpotent (anti-)BRST charges (and the transformations they generate) in the relationship $`(/\theta )(/\overline{\theta })+(/\overline{\theta })(/\theta )=0`$. It should be re-emphasized, however, that all these nice geometrical connections between the BRST formalism and the usual superfield formalism \[9-14\] are confined only to the gauge fields and the (anti-)ghost fields of a BRST invariant Lagrangian density of the $`D`$-dimensional interacting $`p`$-form Abelian gauge theory. The matter fields of an interacting $`D`$-dimensional $`p`$-form Abelian gauge theory remain untouched in the above superfield formalism as far as their nilpotent and anticommuting (anti-)BRST symmetry transformations are concerned. The above constraint due to the horizontality condition (where only $`\stackrel{~}{d}`$ and $`d`$ play important roles) has been generalized to the constraints that emerge from the full use of super ($`\stackrel{~}{d},\stackrel{~}{\delta },\stackrel{~}{\mathrm{\Delta }}`$) and ordinary ($`d,\delta ,\mathrm{\Delta }`$) de Rham cohomological operators (see, e.g., \[16-20\] for details). These complete set of restrictions on the supermanifold lead to the existence of (anti-)BRST, (anti-)co-BRST and a bosonic (which is equal to the anticommutator of the (anti-)BRST and the (anti-)co-BRST) symmetry transformations together for the $`(1+1)`$-dimensional non-interacting 1-form (non-)Abelian gauge theories. In the Lagrangian formulation, the above kind of symmetries have also been shown to exist for the $`(3+1)`$-dimensional free Abelian 2-form gauge theory . There exists a discrete symmetry transformation for the above field theoretical models (in the Lagrangian formulation) which corresponds to the Hodge duality $``$ operation of differential geometry. Thus, the above models do provide a tractable set of field theoretical examples for the Hodge theory. It is worthwhile to pinpoint, however, that even the above new attempts of the superfield formalism (with the full set of cohomological operators) do not shed any light on the nilpotent symmetry transformations associated with the matter fields. In a set of recent papers \[23-27\], the above usual superfield formalism (with the theoretical arsenal of horizontality condition and its generalizations) has been augmented to include the invariance of the conserved currents and/or charges on the supermanifolds. The latter constraints, on the supermanifolds, lead to the derivation of the nilpotent and anticommuting (anti-)BRST symmetry transformations for the matter fields of the interacting four dimensional 1-form (non-)Abelian gauge theories. We christen this extended version of the superfield formalism as the augmented superfield formulation applied to the four dimensional interacting 1-form (non-)Abelian gauge theories described by the corresponding (anti-)BRST invariant Lagrangian densities. It is worth emphasizing that, in the framework of augmented superfield formalism, all the geometrical interpretations, listed in the previous paragraph, remain intact. As a consequence, there is a very nice mutual consistency and complementarity between the old constraint (i.e. the horizontality condition) and new constraint(s) on the supermanifolds. We do obtain, as a bonus and by-product, all the nilpotent (anti-)BRST transformations for all the fields (i.e. gauge fields, (anti-)ghost fields and matter fields) of an interacting 1-form (non-)Abelian gauge theory. The purpose of the present paper is to derive the nilpotent (anti-)BRST transformations for all the fields, present in the description of a free spinning relativistic particle (moving on a super world-line) in the framework of augmented superfield formulation \[23-27\]. Our present endeavour is essential primarily on four counts. First and foremost, this formalism is being applied to a supersymmetric system for the first time. It is worth pointing out that its non-supersymmetric counterpart (i.e. the system of a free scalar relativistic particle) has already been discussed in the framework of augmented superfield formulation in our earlier work . Second, to check the mutual consistency and complementarity between (i) the horizontality condition, and (ii) the invariance of conserved quantities on the supermanifold for this physical system. These were found to be true in the case of (i) a free scalar relativistic particle , (ii) the interacting (non-)Abelian gauge theories in two $`(1+1)`$-dimensions (2D) , (iii) the interacting (non-)Abelian gauge theories in four $`(3+1)`$-dimensions (4D) of spacetime . Third, to generalize our earlier works \[23-27\], which were connected only with the gauge symmetries and reparametrization symmetries, to the case where the supergauge symmetry also exists for the present system under discussion. Finally, to tap the potential and power of the above restrictions in the derivation of the nilpotent symmetries for the case of a new system where the fermionic as well as bosonic (i) the gauge fields (i.e. $`\chi ,e)`$, and (ii) the (anti-)ghost fields (i.e. $`(\overline{c})c`$) and $`(\overline{\beta })\beta `$) do exist in the Lagrangian description of this supersymmetric system (cf. (2.7) below). The contents of our present paper are organized as follows. In Sec. 2, we very clearly discuss the essentials of the reparametrization, gauge and supergauge symmetry transformations for the spinning massive relativistic particle in the Lagrangian formulation. Two sets of anticommuting BRST symmetry transformations, that exist for the above system under a very specific limit, are also discussed in this section. Sections 3 and 4 are the centrals of our paper. Sec. 3 is devoted to the derivation of the nilpotent (anti-)BRST symmetry transformations (corresponding to the gauge symmetry transformations) in the framework of augmented superfield formalism. In the forthcoming section (i.e. Sec. 4), for the first time, we extend the idea of augmented superfield formalism to obtain the (anti-)BRST symmetry transformations (corresponding to the supergauge symmetry transformations) that exist for the spinning relativistic particle. Finally, in Sec. 5, we make some concluding remarks and point out a few future directions for further investigations. 2 Preliminary: Nilpotent BRST Symmetries Let us begin with the various equivalent forms of the reparametrization invariant Lagrangians for the description of a free massive spinning relativistic particle moving on a super world-line that is embedded in a $`D`$-dimensional flat Minkowski target spacetime manifold. These, triplets of appropriate Lagrangians, are : $$\begin{array}{ccc}& & L_0^{(m)}=m[(\dot{x}_\mu +i\chi \psi _\mu )^2]^{1/2}+\frac{i}{2}\left(\psi _\mu \dot{\psi }^\mu \psi _5\dot{\psi }_5\right)i\chi \psi _5m,\hfill \\ & & L_f^{(m)}=p_\mu \dot{x}^\mu \frac{1}{2}e(p^2m^2)+\frac{i}{2}\left(\psi _\mu \dot{\psi }^\mu \psi _5\dot{\psi }_5\right)+i\chi \left(\psi _\mu p^\mu \psi _5m\right),\hfill \\ & & L_s^{(m)}=\frac{1}{2}e^1\left(\dot{x}_\mu +i\chi \psi _\mu \right)^2+\frac{1}{2}em^2+\frac{i}{2}\left(\psi _\mu \dot{\psi }^\mu \psi _5\dot{\psi }_5\right)i\chi \psi _5m.\hfill \end{array}$$ $`(2.1)`$ In the above, the mass-shell condition ($`p^2m^2=0`$), the constraint condition $`p\psi m\psi _5=0`$ and the force free (i.e. $`\dot{p}_\mu =0`$) motion of the spinning relativistic particle are some of the key common features for (i) the Lagrangian with the square root $`L_0^{(m)}`$, (ii) the first-order Lagrangian $`L_f^{(m)}`$, and (iii) the second-order Lagrangian $`L_s^{(m)}`$. The constraints $`p^2m^20`$ and $`p\psi m\psi _50`$ in $`L_f^{(m)}`$ are taken care of by the Lagrange multiplier fields $`e(\tau )`$ and $`\chi (\tau )`$ (with $`\chi ^2=0`$) which are (i) the bosonic and fermionic gauge fields of the present system, respectively, and (ii) the analogues of the vierbein and Rarita-Schwinger (gravitino) fields in the language of the supergravity theories. The Lorentz vector fermionic fields $`\psi _\mu (\tau )`$ (with $`\mu =0,1,2\mathrm{}.D1`$) are the superpartner of the target space coordinate variable $`x_\mu (\tau )`$ (with $`\mu =0,1,2\mathrm{}.D1`$) and classically they present spin degrees of freedom. Furthermore, they anticommute with themselves (i.e. $`\psi _\mu \psi _\nu +\psi _\nu \psi _\mu =0`$) and other fermionic field variables (i.e. $`\psi _\mu \psi _5+\psi _5\psi _\mu =0,\psi _\mu \chi +\chi \psi _\mu =0`$) of the system under consideration. The $`\tau `$-independent mass parameter $`m`$ (i.e. the analogue of the cosmological constant term) is introduced in our present system through the anticommuting (i.e. $`\psi _5\chi +\chi \psi _5=0,(\psi _5)^2=1`$, etc.) Lorentz scalar field $`\psi _5(\tau )`$. The momenta $`p_\mu (\tau )`$ (with $`\mu =0,1,2\mathrm{}.D1`$), present in $`L_f^{(m)}`$, are canonically conjugate to the target space coordinate variable $`x^\mu (\tau )`$. It is evident that, except for the mass parameter $`m`$, the rest of the field variables are the functions of monotonically increasing parameter $`\tau `$ that characterizes the trajectory (i.e. the super world-line) of the massive spinning relativistic particle. Here $`\dot{x}^\mu =(dx^\mu /d\tau )=ep^\mu i\chi \psi ^\mu `$, $`\dot{\psi }_\mu =(d\psi _\mu /d\tau )=\chi p_\mu ,\dot{\psi }_5=(d\psi _5/d\tau )=\chi m`$ are the generalized versions of “velocities” of the massive spinning relativistic particle. In what follows, we shall focus on the first-order Lagrangian $`L_f^{(m)}`$ for the discussion of the symmetry properties of the system. This is due to the fact that this Lagrangian is comparatively simpler in the sense that there are no square roots and there are no field variables in the denominator. Furthermore, it is endowed with the maximum number of field variables and, therefore, is interesting from the point of view of theoretical discussions. Under an infinitesimal version of the reparametrization transformations $`\tau \tau ^{}=\tau ϵ(\tau )`$, where $`ϵ(\tau )`$ is an infinitesimal parameter, the field variables of $`L_f^{(m)}`$ transform as $$\begin{array}{ccc}\delta _rx_\mu \hfill & =& ϵ\dot{x}_\mu ,\delta _rp_\mu =ϵ\dot{p}_\mu ,\delta _r\psi _\mu =ϵ\dot{\psi }_\mu ,\hfill \\ \delta _r\psi _5\hfill & =& ϵ\dot{\psi }_5,\delta _r\chi =\frac{d}{d\tau }\left(ϵ\chi \right),\delta _re=\frac{d}{d\tau }\left(ϵe\right).\hfill \end{array}$$ $`(2.2)`$ It should noted that (i) $`\delta _r\mathrm{\Sigma }(\tau )=\mathrm{\Sigma }^{}(\tau )\mathrm{\Sigma }(\tau )`$ for the generic field variable $`\mathrm{\Sigma }=x_\mu ,p_\mu ,e,\psi _\mu ,\psi _5,\chi `$, and (ii) the gauge fields $`e`$ and $`\chi `$ do transform in a similar fashion (and distinctly different from the rest of the field variables). The first- and the second-order Lagrangians are endowed with the first-class constraints $`\mathrm{\Pi }_e0,\mathrm{\Pi }_\chi 0,p^2m^20,p\psi m\psi _50`$ in the language of Dirac’s prescription for the classification of constraints. Here $`\mathrm{\Pi }_e`$ and $`\mathrm{\Pi }_\chi `$ are the canonical conjugate momenta corresponding to the einbein field $`e(\tau )`$ and the fermionic gauge field $`\chi `$, respectively. There are second-class constraints too in the theory but we shall not concentrate on them for our present discussion. The existence of the first-class constraints $`\mathrm{\Pi }_e0`$ and $`p^2m^20`$ on this physical system, generates the following gauge symmetry transformation $`\delta _g`$ for the field variables of the first-order Lagrangian $`L_f^{(m)}`$ (for the description of a spinning relativistic particle): $$\begin{array}{ccc}\delta _gx_\mu \hfill & =& \xi p_\mu ,\delta _gp_\mu =0,\delta _g\psi _\mu =0,\hfill \\ \delta _g\psi _5\hfill & =& 0,\delta _g\chi =0,\delta _ge=\dot{\xi },\hfill \end{array}$$ $`(2.3)`$ where $`\xi (\tau )`$ is an infinitesimal gauge parameter. The pair of fermionic constraints $`\pi _\chi 0`$ and $`p\psi m\psi _50`$, generate the following supergauge symmetry transformations $`\delta _{sg}`$ for the bosonic and fermionic field variables of the first-order Lagrangian $`L_f^{(m)}`$: $$\begin{array}{ccc}\delta _{sg}x_\mu \hfill & =& \kappa \psi _\mu ,\delta _{sg}p_\mu =0,\delta _{sg}\psi _\mu =i\kappa p_\mu ,\hfill \\ \delta _{sg}\psi _5\hfill & =& i\kappa m,\delta _{sg}\chi =i\dot{\kappa },\delta _{sg}e=2\kappa \chi ,\hfill \end{array}$$ $`(2.4)`$ where $`\kappa (\tau )`$ is an infinitesimal fermionic (i.e. $`\kappa ^2=0`$) supergauge transformation parameter. The above infinitesimal transformations are symmetry transformations because: $$\begin{array}{ccc}\delta _rL_f^{(m)}\hfill & =& \frac{d}{d\tau }\left[ϵL_f^{(m)}\right],\delta _gL_f^{(m)}=\frac{d}{d\tau }\left[\frac{\xi }{2}\left(p^2+m^2\right)\right],\hfill \\ \delta _{sg}L_f^{(m)}\hfill & =& \frac{d}{d\tau }\left[\frac{\kappa }{2}\left(p\psi +m\psi _5\right)\right].\hfill \end{array}$$ $`(2.5)`$ It is straightforward to check that, for the generic field variable $`\mathrm{\Sigma }`$, we have $`(\delta _gi\delta _{sg})\mathrm{\Sigma }=\delta _r\mathrm{\Sigma }`$ with the identifications $`\xi =eϵ`$ and $`\kappa =\chi ϵ`$ and validity of the on-shell conditions (i.e. $`\dot{p}_\mu =0,\dot{\psi }_\mu =\chi p_\mu ,\dot{\psi }_5=\chi m,\dot{x}_\mu =ep_\mu i\chi \psi _\mu ,p^2=m^2,p\psi =m\psi _5`$). The gauge- and supergauge symmetry transformations (2.3) and (2.4) can be combined together and generalized to the nilpotent BRST symmetry transformations. The usual trick of the BRST prescription could be exploited here to express the gauge- and supergauge parameters $`\xi =\eta c`$ and $`\kappa =\eta \beta `$ in terms of the fermionic ($`c^2=0`$) and bosonic ($`\beta ^20`$) ghost fields and $`\eta `$. It will be noted that $`\eta `$ is the spacetime independent anticommuting (i.e. $`\eta c+c\eta =0`$, etc.) parameter which is required to maintain the bosonic nature of $`\xi `$ (in $`\xi =\eta c`$) and the fermionic nature of $`\kappa `$ (in $`\kappa =\eta \beta `$). The ensuing nilpotent ($`(s_b^{(0)})^2=0`$) BRST transformations , for the spinning relativistic particle, are <sup>§</sup><sup>§</sup>§We follow here the notations and conventions adopted in . In its full blaze of glory, the true nilpotent (anti-)BRST transformations $`\delta _{(A)B}`$ are the product of an (anticommuting) spacetime independent parameter $`\eta `$ and the nilpotent transformations $`s_{(a)b}`$. It is clear that $`\eta `$ commutes with all the bosonic (even) fields of the theory and anticommutes with fermionic (odd) fields (i.e. $`\eta c+c\eta =0,\eta \overline{c}+\overline{c}\eta =0`$, etc.). $$\begin{array}{ccc}s_b^{(0)}x_\mu \hfill & =& cp_\mu +\beta \psi _\mu ,s_b^{(0)}c=i\beta ^2,s_b^{(0)}p_\mu =0,s_b^{(0)}\psi _\mu =i\beta p_\mu ,\hfill \\ s_b^{(0)}\overline{c}\hfill & =& ib,s_b^{(0)}b=0,s_b^{(0)}e=\dot{c}+2\beta \chi ,s_b^{(0)}\chi =i\dot{\beta },\hfill \\ s_b^{(0)}\psi _5\hfill & =& i\beta m,s_b^{(0)}\beta =0,s_b^{(0)}\overline{\beta }=i\gamma ,s_b^{(0)}\gamma =0.\hfill \end{array}$$ $`(2.6)`$ The above off-shell nilpotent ($`(s_b^{(0)})^2=0`$) transformations are the symmetry transformations for the system because the following Lagrangian (which is the generalization of $`L_f^{(m)}`$): $$\begin{array}{ccc}L_b^{(m)}\hfill & =& p\dot{x}\frac{1}{2}e(p^2m^2)+\frac{i}{2}\left(\psi \dot{\psi }\psi _5\dot{\psi }_5\right)+i\chi \left(p\psi \psi _5m\right)\hfill \\ & +& b\dot{e}+\gamma \dot{\chi }+\frac{1}{2}b^2i\dot{\overline{c}}\left(\dot{c}+2\chi \beta \right)\dot{\overline{\beta }}\dot{\beta },\hfill \end{array}$$ $`(2.7)`$ transforms to a total derivative under (2.6), namely; $$\begin{array}{ccc}s_b^{(0)}L_b^{(m)}=\frac{d}{d\tau }\left[\frac{1}{2}c(p^2+m^2)+\frac{1}{2}\beta (p\psi +m\psi _5)+b(\dot{c}+2\beta \chi )i\gamma \dot{\beta }\right].\hfill & & \end{array}$$ $`(2.8)`$ A few comments are in order now. First, the first-order Lagrangian $`L_f^{(m)}`$ has been extended to include the gauge-fixing term and the Faddeev-Popov ghost terms (constructed by the bosonic as well as the fermionic (anti-)ghost fields) in $`L_b^{(m)}`$. Second, the bosonic auxiliary field $`b`$ and the fermionic auxiliary field $`\gamma `$ (with $`\gamma ^2=0,\gamma \chi +\chi \gamma =0,c\gamma +\gamma c=0`$, etc.) are the Nakanishi-Lautrup fields. Third, the fermionic (i.e. $`\overline{c}^2=0,c\overline{c}+\overline{c}c=0,`$ etc.) anti-ghost field $`\overline{c}`$ and the bosonic (i.e. $`\overline{\beta }^20,\beta \overline{\beta }=\overline{\beta }\beta `$, etc.) anti-ghost field $`\overline{\beta }`$ are required in the theory to have a precise nilpotent BRST symmetry for the system under consideration. Fourth, the above nilpotent transformations (2.6) are generated by the conserved ($`\dot{Q_b^{(0)}}=0`$) and nilpotent (i.e. $`(Q_b^{(0)})^2=0`$) BRST charge $`Q_b^{(0)}`$ as given below: $$\begin{array}{ccc}Q_b^{(0)}=\frac{c}{2}\left(p^2m^2\right)+\beta \left(p\psi m\psi _5\right)+b\left(\dot{c}+2\beta \chi \right)+\dot{\overline{c}}\beta ^2i\gamma \dot{\beta }.\hfill & & \end{array}$$ $`(2.9)`$ Fifth, the nilpotent ($`(s_{ab}^{(0)})^2=0`$) anti-BRST symmetry transformations $`s_{ab}^{(0)}`$ and corresponding generator $`Q_{ab}^{(0)}`$ can be computed from (2.6) and (2.9) by the substitutions: $`c\overline{c}`$ and $`\beta \overline{\beta }`$. Sixth, the above generators and corresponding symmetries obey the property of anticommutativity (i.e. $`s_b^{(0)}s_{ab}^{(0)}+s_{ab}^{(0)}s_b^{(0)}=0,Q_b^{(0)}Q_{ab}^{(0)}+Q_{ab}^{(0)}Q_b^{(0)}=0`$). Finally, the conservation of the BRST charge $`Q_b^{(0)}`$ can be proven by exploiting the equations of motion $$\begin{array}{ccc}& & \dot{p}_\mu =0,\dot{x}_\mu =ep_\mu i\chi \psi _\mu ,\dot{\psi }_\mu =\chi p_\mu ,\dot{\psi }_5=\chi m,\dot{\chi }=0,\hfill \\ & & b=\dot{e},\dot{b}=\frac{1}{2}(p^2m^2),\ddot{\beta }=0,\ddot{\overline{\beta }}=2i\dot{\overline{c}}\chi ,\ddot{\overline{c}}=0,\hfill \\ & & \ddot{c}+2\dot{\beta }\chi +2\beta \dot{\chi }=0,\dot{\gamma }+2i\dot{\overline{c}}\beta +i(p\psi m\psi _5)=0,\hfill \end{array}$$ $`(2.10)`$ derived from the BRST invariant Lagrangian $`L_b^{(m)}`$ of (2.7). For our further discussion, we deal with the limiting cases of (2.6) and (2.7) so that we can study the BRST transformations corresponding to the gauge transformations (2.3) and the supergauge transformations (2.4), separately and independently. It is evident that $`\beta 0,\overline{\beta }0,\gamma 0`$ in (2.6) leads to the nilpotent (($`s_b^{(1)})^2=0`$) BRST transformations $`s_b^{(1)}`$, corresponding to the gauge transformations (2.3), as $$\begin{array}{ccc}s_b^{(1)}x_\mu \hfill & =& cp_\mu ,s_b^{(1)}p_\mu =0,s_b^{(1)}c=0,s_b^{(1)}\psi _\mu =0,\hfill \\ s_b^{(1)}\psi _5\hfill & =& 0,s_b^{(1)}\overline{c}=ib,s_b^{(1)}b=0,s_b^{(1)}\chi =0,s_b^{(1)}e=\dot{c},\hfill \end{array}$$ $`(2.11)`$ which are the symmetry transformations for the Lagrangian $$\begin{array}{ccc}L_b^{(1)}\hfill & =& p\dot{x}\frac{1}{2}e(p^2m^2)+\frac{i}{2}\left(\psi \dot{\psi }\psi _5\dot{\psi }_5\right)\hfill \\ & +& i\chi \left(p\psi \psi _5m\right)+b\dot{e}+\frac{1}{2}b^2i\dot{\overline{c}}\dot{c},\hfill \end{array}$$ $`(2.12)`$ obtained from (2.7) under the above specific limits (i.e. $`(\beta ,\overline{\beta },\gamma )0`$). In another limiting case (i.e. $`c0,\overline{c}0,b0`$) of (2.6), we obtain the following non-nilpotent ($`(s_b^{(2)})^20`$) BRST transformations, corresponding to the supergauge transformations (2.4), as $$\begin{array}{ccc}s_b^{(2)}x_\mu \hfill & =& \beta \psi _\mu ,s_b^{(2)}p_\mu =0,s_b^{(2)}\beta =0,s_b^{(2)}\psi _\mu =i\beta p_\mu ,\hfill \\ s_b^{(2)}\psi _5\hfill & =& i\beta m,s_b^{(2)}\chi =i\dot{\beta },s_b^{(2)}\overline{\beta }=i\gamma ,s_b^{(2)}\gamma =0,s_b^{(2)}e=2\beta \chi ,\hfill \end{array}$$ $`(2.13)`$ that are found to be the symmetry transformations for the Lagrangian $$\begin{array}{ccc}L_b^{(2)}\hfill & =& p\dot{x}\frac{1}{2}e(p^2m^2)+\frac{i}{2}\left(\psi \dot{\psi }\psi _5\dot{\psi }_5\right)\hfill \\ & +& i\chi \left(p\psi \psi _5m\right)+\gamma \dot{\chi }\dot{\overline{\beta }}\dot{\beta },\hfill \end{array}$$ $`(2.14)`$ derived from (2.7) under the above limits: $`(c,\overline{c},b)0`$. It will be noted that (i) the anti-BRST versions of (2.11) and (2.13) can be obtained by the substitutions $`c\overline{c}`$ and $`\beta \overline{\beta }`$, respectively. (ii) In a similar fashion, the generators $`Q_b^{(1,2)}`$ for $`s_b^{(1,2)}`$, can be derived from (2.9) by taking into account the above limiting cases. (iii) The equations of motion for the Lagrangians (2.12) and (2.14) can be derived from (2.10) under the limits cited above. We wrap up this section with a couple of comments on the nilpotency property of the BRST transformations $`s_b^{(2)}`$ in (2.13) that correspond to the supergauge transformations in (2.4). First, it is straightforward to note that, under the restriction $`\beta ^2=0`$, one can restore the nilpotency (i.e. $`(s_b^{(2)})^2=0`$) for the transformations $`s_b^{(2)}`$. This aspect of $`\beta `$ can be fulfilled if this bosonic ghost field $`\beta `$ is taken to be a composite (i.e. $`\beta c_1c_2`$) of a couple of fermionic ($`c_1^2=c_2^2=0,c_1c_2+c_2c_1=0`$) ghost fields $`c_1`$ and $`c_2`$ . Second, if this condition (i.e. $`\beta ^2=0`$) is true, it turns out that the two BRST symmetry transformations $`s_b^{(1,2)}`$ really decouple from each-other in the sense that $`\{s_b^{(1)},s_b^{(2)}\}=0`$. In such a situation, they can distinctly be separate from each-other as well as from $`s_b`$ in (2.6) (with $`(s_b^{(1,2)})^2=0,\{s_b^{(1)},s_b^{(2)}\}=0`$) . It is, ultimately, interesting to note that all the nilpotent symmetry transformations $`s_r^{(i)}`$ (with $`i=0,1,2`$), for the generic field $`\mathrm{\Sigma }`$ of the system, can be succinctly expressed in terms of the conserved and nilpotent charges $`Q_r^{(i)}`$ (with $`i=0,1,2`$) as $$\begin{array}{ccc}s_r^{(i)}\mathrm{\Sigma }=i[\mathrm{\Sigma },Q_r^{(i)}]_\pm ,r=b,ab,\mathrm{\Sigma }=x_\mu ,p_\mu ,b,e,c,\overline{c},\chi ,\gamma ,\beta ,\overline{\beta },\psi _\mu ,\psi _5.\hfill & & \end{array}$$ $`(2.15)`$ The $`\pm `$ signs, present as the subscripts on the above square brackets, stand for the brackets to be (anti-)commutators for the generic field $`\mathrm{\Sigma }`$ being (fermionic)bosonic in nature. 3 Gauge BRST Symmetries: Augmented Superfield Approach To derive the nilpotent ($`(s_{(a)b}^{(1)})^2=0`$) (anti-)BRST transformations $`s_{(a)b}^{(1)}`$ (cf. (2.11)) for the gauge (einbein) field $`e(\tau )`$ and the (anti-)ghost fields $`(\overline{c})c`$ in the superfield formalism (where the horizontality condition plays a decisive role), we begin with a general three ($`1+2`$)-dimensional supermanifold parametrized by the superspace coordinates $`Z=(\tau ,\theta ,\overline{\theta })`$ where $`\tau `$ is an even (bosonic) coordinate and $`\theta `$ and $`\overline{\theta }`$ are the two odd (Grassmannian) coordinates (with $`\theta ^2=\overline{\theta }^2=0,\theta \overline{\theta }+\overline{\theta }\theta =0)`$. On this supermanifold, one can define a 1-form supervector superfield $`\stackrel{~}{V}=dZ(\stackrel{~}{A})`$ with $`\stackrel{~}{A}(\tau ,\theta ,\overline{\theta })=(E(\tau ,\theta ,\overline{\theta }),F(\tau ,\theta ,\overline{\theta }),\overline{F}(\tau ,\theta ,\overline{\theta }))`$ as the component multiplet superfields. The superfields $`E,F,\overline{F}`$ can be expanded in terms of the basic fields ($`e,c,\overline{c}`$) and auxiliary field ($`b`$) along with some extra secondary fields (i.e. $`f,\overline{f},B,g,\overline{g},s,\overline{s},\overline{b}`$), as given below (see, e.g., ): $$\begin{array}{ccc}E(\tau ,\theta ,\overline{\theta })\hfill & =& e(\tau )+\theta \overline{f}(\tau )+\overline{\theta }f(\tau )+i\theta \overline{\theta }B(\tau ),\hfill \\ F(\tau ,\theta ,\overline{\theta })\hfill & =& c(\tau )+i\theta \overline{b}(\tau )+i\overline{\theta }g(\tau )+i\theta \overline{\theta }s(\tau ),\hfill \\ \overline{F}(\tau ,\theta ,\overline{\theta })\hfill & =& \overline{c}(\tau )+i\theta \overline{g}(\tau )+i\overline{\theta }b(\tau )+i\theta \overline{\theta }\overline{s}(\tau ).\hfill \end{array}$$ $`(3.1)`$ It is straightforward to note that the local fields $`f(\tau ),\overline{f}(\tau ),c(\tau ),\overline{c}(\tau ),s(\tau ),\overline{s}(\tau )`$ on the r.h.s. are fermionic (anti-commuting) in nature and the bosonic (commuting) local fields in (3.1) are: $`e(\tau ),B(\tau ),g(\tau ),\overline{g}(\tau ),b(\tau ),\overline{b}(\tau )`$. It is evident that, in the above expansion, the bosonic- and fermionic degrees of freedom match. This requirement is essential for the validity and sanctity of any arbitrary supersymmetric theory in the superfield formulation. In fact, all the secondary fields will be expressed in terms of basic fields due to the restrictions emerging from the application of the horizontality condition (see, e.g., ) $$\begin{array}{ccc}\stackrel{~}{d}\stackrel{~}{V}=dA=0,d=d\tau _\tau ,A=d\tau e(\tau ),d^2=0,\hfill & & \end{array}$$ $`(3.2)`$ where the super exterior derivative $`\stackrel{~}{d}`$ and the super connection 1-form $`\stackrel{~}{V}`$ are defined as $$\begin{array}{ccc}\stackrel{~}{d}\hfill & =& d\tau _\tau +d\theta _\theta +d\overline{\theta }_{\overline{\theta }},\hfill \\ \stackrel{~}{V}\hfill & =& d\tau E(\tau ,\theta ,\overline{\theta })+d\theta \overline{F}(\tau ,\theta ,\overline{\theta })+d\overline{\theta }F(\tau ,\theta ,\overline{\theta }).\hfill \end{array}$$ $`(3.3)`$ It will be noted that super 1-form connection $`\stackrel{~}{V}`$ is overall bosonic in nature because of the fact that the superfields $`E`$ and ($`F,\overline{F}`$) are bosonic ($`E^20`$) and fermionic ($`F^2=\overline{F}^2=0`$), respectively. They have been combined together with $`d\tau `$ and ($`d\theta ,d\overline{\theta }`$) in such a fashion that $`\stackrel{~}{V}`$ becomes bosonic. We expand $`\stackrel{~}{d}\stackrel{~}{V}`$, present in the l.h.s. of (3.2), as $$\begin{array}{ccc}\stackrel{~}{d}\stackrel{~}{V}\hfill & =& (d\tau d\theta )(_\tau \overline{F}_\theta E)(d\theta d\theta )(_\theta \overline{F})+(d\tau d\overline{\theta })(_\tau F_{\overline{\theta }}E),\hfill \\ & & (d\theta d\overline{\theta })(_\theta F+_{\overline{\theta }}\overline{F})(d\overline{\theta }d\overline{\theta })(_{\overline{\theta }}F).\hfill \end{array}$$ $`(3.4)`$ Ultimately, the application of the horizontality condition ($`\stackrel{~}{d}\stackrel{~}{V}=dA=0`$) yields $$\begin{array}{ccc}f(\tau )\hfill & =& \dot{c}(\tau ),\overline{f}(\tau )=\dot{\overline{c}}(\tau ),s(x)=\overline{s}(x)=0,\hfill \\ B(\tau )\hfill & =& \dot{b}(\tau ),b(\tau )+\overline{b}(\tau )=0,g(\tau )=\overline{g}(\tau )=0.\hfill \end{array}$$ $`(3.5)`$ It may be emphasized that the gauge (einbein) field $`e(\tau )`$ is a scalar potential depending only on a single variable parameter $`\tau `$. This is why the curvature is zero (i.e. $`dA=0`$) because $`d\tau d\tau =0`$ It is interesting to point out that, unlike the above case, for the 1-form ($`A=dx^\mu A_\mu `$) Abelian gauge theory, where the gauge field is a vector potential $`A_\mu (x)`$, the 2-form curvature $`dA=\frac{1}{2}(dx^\mu dx^\nu )F_{\mu \nu }`$ is not equal to zero and it defines the field strength tensor $`F_{\mu \nu }=_\mu A_\nu _\nu A_\mu `$ for the Abelian gauge theory. For the 1-form non-Abelian gauge theory, the Maurer-Cartan equation $`F=dA+AA`$ defines the 2-form $`F`$ which, in turn, leads to the derivation of the corresponding group valued field strength tensor $`F_{\mu \nu }`$.. The insertion of all the above values (cf. (3.5)) in the expansion (3.1) leads to the derivation of the (anti-)BRST symmetry transformations for the gauge- and (anti-)ghost fields of the theory. This statement can be expressed, in an explicit form, as given below: $$\begin{array}{ccc}E(\tau ,\theta ,\overline{\theta })\hfill & =& e(\tau )+\theta \dot{\overline{c}}(\tau )+\overline{\theta }\dot{c}(\tau ))+i\theta \overline{\theta }\dot{b}(\tau ),\hfill \\ F(\tau ,\theta ,\overline{\theta })\hfill & =& c(\tau )i\theta b(\tau ),\hfill \\ \overline{F}(\tau ,\theta ,\overline{\theta })\hfill & =& \overline{c}(\tau )+i\overline{\theta }b(\tau ).\hfill \end{array}$$ $`(3.6)`$ In addition, this exercise provides the physical interpretation for the (anti-)BRST charges $`Q_{(a)b}^{(1)}`$ as the generators (cf. (2.15)) of translations (i.e. $`\text{Lim}_{\overline{\theta }0}(/\theta ),\text{Lim}_{\theta 0}(/\overline{\theta })`$) along the Grassmannian directions of the supermanifold. Both these observations can be succinctly expressed, in a combined way, by re-writing the super expansion (3.1) as It is worthwhile to note that the anti-BRST transformations $`s_{ab}^{(1)}`$ for the system, described by the Lagrangian in (2.12), are: $`s_{ab}^{(1)}x_\mu =\overline{c}p_\mu ,s_{ab}^{(1)}\overline{c}=0,s_{ab}^{(1)}p_\mu =0,s_{ab}^{(1)}c=ib,s_{ab}^{(1)}b=0,s_{ab}^{(1)}\chi =0,s_{ab}^{(1)}\psi _5=0,s_{ab}^{(1)}\psi _\mu =0,s_{ab}^{(1)}e=\dot{\overline{c}}`$. The key point, that should be emphasized, is the minus sign in : $`s_{ab}^{(1)}c=ib`$. $$\begin{array}{ccc}E(\tau ,\theta ,\overline{\theta })\hfill & =& e(\tau )+\theta (s_{ab}^{(1)}e(\tau ))+\overline{\theta }(s_b^{(1)}e(\tau ))+\theta \overline{\theta }(s_b^{(1)}s_{ab}^{(1)}e(\tau )),\hfill \\ F(\tau ,\theta ,\overline{\theta })\hfill & =& c(\tau )+\theta (s_{ab}^{(1)}c(\tau ))+\overline{\theta }(s_b^{(1)}c(\tau ))+\theta \overline{\theta }(s_b^{(1)}s_{ab}^{(1)}(\tau )),\hfill \\ \overline{F}(\tau ,\theta ,\overline{\theta })\hfill & =& \overline{c}(\tau )+\theta (s_{ab}^{(1)}\overline{c}(\tau ))+\overline{\theta }(s_b^{(1)}\overline{c}(\tau ))+\theta \overline{\theta }(s_b^{(1)}s_{ab}^{(1)}\overline{c}(\tau )).\hfill \end{array}$$ $`(3.7)`$ It should be noted that the third and fourth terms of the expansion for $`F`$ and the second and fourth terms of the expansion for $`\overline{F}`$ are zero because ($`s_b^{(1)}c=0,s_{ab}^{(1)}\overline{c}=0`$). Let us concentrate on the derivation of the nilpotent (i.e. $`(s_{(a)b}^{(1)})^2=0`$) (anti-)BRST symmetry transformations $`s_{(a)b}^{(1)}`$ for the Lorentz scalar fields $`\chi (\tau ),\psi _5(\tau )`$ and the Lorentz vector target fields $`(x_\mu (\tau ),\psi _\mu (\tau ),p_\mu (\tau ))`$. In the derivation of the nilpotent transformations for these fields, under the framework of augmented superfield formalism, it is the invariance of the conserved quantities on the supermanifold that plays a key role. However, at times, one has to tap the inputs from the super expansions (3.6), derived after the application of the horizontality condition, for the precise derivations of the nilpotent transformations. Thus, to be very precise, it is the interplay of the horizontality condition and the invariance of the conserved charges that enables us to derive the nilpotent (anti-)BRST transformations. To justify this assertion, first of all, we start off with the super expansion of the superfields $`(X^\mu ,P_\mu )(\tau ,\theta ,\overline{\theta })`$), corresponding to the ordinary target variables $`(x^\mu ,p_\mu )(\tau )`$ (that specify the Minkowski cotangent manifold and are present in the first-order Lagrangian $`L_f^{(m)}`$), as $$\begin{array}{ccc}X_\mu (\tau ,\theta ,\overline{\theta })\hfill & =& x_\mu (\tau )+\theta \overline{R}_\mu (\tau )+\overline{\theta }R_\mu (\tau )+i\theta \overline{\theta }S_\mu (\tau ),\hfill \\ P_\mu (\tau ,\theta ,\overline{\theta })\hfill & =& p_\mu (\tau )+\theta \overline{F}_\mu (\tau )+\overline{\theta }F_\mu (\tau )+i\theta \overline{\theta }T_\mu (\tau ).\hfill \end{array}$$ $`(3.8)`$ It is evident that, in the limit $`(\theta ,\overline{\theta })0`$, we get back the canonically conjugate target space variables $`(x^\mu (\tau ),p_\mu (\tau ))`$ of the first-order Lagrangian $`L_f^{(m)}`$ in (2.1). Furthermore, the number of bosonic fields ($`x_\mu ,p_\mu ,S_\mu ,T_\mu )`$ do match with the fermionic fields $`(F_\mu ,\overline{F}_\mu ,R_\mu ,\overline{R}_\mu )`$ so that the above expansion becomes consistent with the basic tenets of supersymmetry. All the component fields on the r.h.s. of the expansion (3.8) are functions of the monotonically increasing parameter $`\tau `$ of the world-line. As emphasized in Section 2, three most decisive features of the free relativistic particle are (i) $`\dot{p}_\mu =0`$, (ii) $`p\psi m\psi _5=0`$, and (ii) $`p^2m^2=0`$. To be very specific, it can be seen that the conserved gauge charge $`Q_g=\frac{1}{2}(p^2m^2)`$ couples to the ‘gauge’ (einbein) field $`e(\tau )`$ in the Lagrangian $`L_f^{(m)}`$ to maintain the local gauge invariance <sup>\**</sup><sup>\**</sup>\**Exactly the same kind of gauge coupling exists between the Dirac fields for the fermions (electrons, positrons, quarks, etc.) and the gauge boson field of the interacting 1-form (non-)Abelian gauge theories where the matter conserved current $`J_\mu =\overline{\psi }\gamma _\mu \psi `$, constructed by the Dirac fields, couples to the gauge field $`A_\mu `$ of the (non-)Abelian gauge theories to maintain the local gauge invariance (see, e.g., ). under the transformations (2.3). For the BRST invariant Lagrangian (2.12), the same kind of coupling exists for the local BRST invariance to be maintained in the theory. The invariance of the mass-shell condition $`(p^2m^2=0`$) (i.e. a conserved and gauge invariant quantity) as well as the conservation of the gauge invariant momenta ($`\dot{p}_\mu =0`$) on the supermanifold, namely; $$\begin{array}{ccc}P_\mu (\tau ,\theta ,\overline{\theta })P^\mu (\tau ,\theta ,\overline{\theta })m^2=p_\mu (\tau )p^\mu (\tau )m^2,\dot{P}_\mu (\tau ,\theta ,\overline{\theta })=\dot{p}_\mu (\tau ),\hfill & & \end{array}$$ $`(3.9)`$ imply the following restrictions: $$\begin{array}{ccc}F_\mu (\tau )=\overline{F}_\mu (\tau )=T_\mu (\tau )=0,\text{and}P_\mu (\tau ,\theta ,\overline{\theta })=p_\mu (\tau ).\hfill & & \end{array}$$ $`(3.10)`$ In other words, the invariance of the mass-shell condition as well as the conserved momenta on the supermanifold enforces $`P_\mu (\tau ,\theta ,\overline{\theta })`$ to be independent of the Grassmannian variables $`\theta `$ and $`\overline{\theta }`$. To be consistent with our earlier interpretations for the (anti-)BRST charges, in the language of translation generators along the Grassmannian directions $`(\theta )\overline{\theta }`$ of the supermanifold, it can be seen that the above equation can be re-expressed as $$\begin{array}{ccc}P_\mu (\tau ,\theta ,\overline{\theta })=p_\mu (\tau )+\theta (s_{ab}^{(1)}p_\mu (\tau ))+\overline{\theta }(s_b^{(1)}p_\mu (\tau ))+\theta \overline{\theta }(s_b^{(1)}s_{ab}^{(1)}p_\mu (\tau )).\hfill & & \end{array}$$ $`(3.11)`$ The above equation, vis-à-vis (3.10), makes it clear that $`s_b^{(1)}p_\mu (\tau )=0`$ and $`s_{ab}^{(1)}p_\mu (\tau )=0`$. Before we shall derive the nilpotent (anti-)BRST transformations for $`x_\mu (\tau )`$, it is useful to compute these transformations for the fermionic gauge field $`\chi (\tau )`$ and other fields $`\psi _5(\tau )`$ as well as $`\psi _\mu (\tau )`$. To this end in mind, let us have the super expansions for the superfields, corresponding to these fields, as follows $$\begin{array}{ccc}K(\tau ,\theta ,\overline{\theta })\hfill & =& \chi (\tau )+\theta \overline{b}_1(\tau )+\overline{\theta }b_1(\tau )+\theta \overline{\theta }f_1(\tau ),\hfill \\ \mathrm{\Psi }_5(\tau ,\theta ,\overline{\theta })\hfill & =& \psi _5(\tau )+\theta \overline{B}_5(\tau )+\overline{\theta }B_5(\tau )+\theta \overline{\theta }f_5(\tau ),\hfill \\ \mathrm{\Psi }_\mu (\tau ,\theta ,\overline{\theta })\hfill & =& \psi _\mu (\tau )+\theta \overline{b}_\mu (\tau )+\overline{\theta }b_\mu (\tau )+\theta \overline{\theta }f_\mu (\tau ).\hfill \end{array}$$ $`(3.12)`$ It will be noted that, in the limit ($`\theta ,\overline{\theta })0`$, we do obtain the usual local fields $`\chi (\tau ),\psi _5(\tau )`$ and $`\psi _\mu (\tau )`$ and the fermionic ($`\chi ,\psi _5,\psi _\mu ,f_\mu ,f_1,f_5`$) and bosonic ($`b_1,\overline{b}_1,B_5,\overline{B}_5,b_\mu ,\overline{b}_\mu `$) degrees of freedom do match in the above expansion. Let us focus on the conserved quantities $`\dot{\chi }=0,\ddot{\psi }_5=\dot{\chi }m=0,\ddot{\psi }_\mu =\dot{\chi }p_\mu +\chi \dot{p}_\mu =0`$ (cf. (2.10)). The invariance of $`\dot{\chi }(\tau )=0`$ on the supermanifold <sup>††</sup><sup>††</sup>††This condition, in a more sophisticated language, is just the gauge choice for the system. leads to the following consequences: $$\begin{array}{ccc}\dot{K}(\tau ,\theta ,\overline{\theta })=\dot{\chi }(\tau )=0\dot{b}_1=\dot{\overline{b}_1}=\dot{f}_1=0.\hfill & & \end{array}$$ $`(3.13)`$ One of the solutions is $`b_1=\overline{b}_1=f_1=C`$ where $`C`$ is a $`\tau `$-independent constant. Geometrically, this amounts to the shift of the superfield $`K(\tau ,\theta ,\overline{\theta })`$ along $`\theta `$\- and $`\overline{\theta }`$-directions by a constant value $`C`$. One can choose the $`\theta `$ and $`\overline{\theta }`$ axes on the supermanifold in such a manner that this constant $`C`$ is zero. Thus, ultimately, we obtain $`b_1=\overline{b}_1=f_1=0`$. Interpreted in the light of (3.6), (3.7) and (3.11), this shows that $`s_{(a)b}^{(1)}\chi =0`$ (i.e. $`K(\tau ,\theta ,\overline{\theta })=\chi (\tau )`$). As a side remark, it is worthwhile to mention that the above explicit equality (3.13) can be re-expressed as the equality of the conserved quantities only. In other words, the restriction $`K(\tau ,\theta ,\overline{\theta })=\chi (\tau )`$, directly implies that $`b_1=\overline{b}_1=f_1=0`$. For the (non-)Abelian gauge theories, such kind of equality has been taken into account \[24-26\] where only the expression for the conserved quantity has been equated on the supermanifold. In view of the above, it can be seen that the following invariances of the conserved quantities (cf. (2.10) for details) on the supermanifold: $$\begin{array}{ccc}& & \dot{\mathrm{\Psi }}_5(\tau ,\theta ,\overline{\theta })K(\tau ,\theta ,\overline{\theta })m=\dot{\psi }_5(\tau )\chi (\tau )m,\hfill \\ & & \dot{\mathrm{\Psi }}_\mu (\tau ,\theta ,\overline{\theta })K(\tau ,\theta ,\overline{\theta })P_\mu (\tau ,\theta ,\overline{\theta })=\dot{\psi }_\mu (\tau )\chi (\tau )p_\mu (\tau ),\hfill \end{array}$$ $`(3.14)`$ imply the following restrictions for the expansion in (3.12), namely; $$\begin{array}{ccc}\overline{B}_5=B_5=f_5=0,\overline{b}_\mu =b_\mu =f_\mu =0.\hfill & & \end{array}$$ $`(3.15)`$ In the above derivation, we have exploited the results of (3.10) (i.e. $`P_\mu (\tau ,\theta ,\overline{\theta })=p_\mu (\tau )`$) and (3.13) (i.e. $`K(\tau ,\theta ,\overline{\theta })=\chi (\tau ))`$. Insertions of (3.15) into (3.12) imply that $`s_{(a)b}^{(1)}\psi _5=0`$ (i.e. $`\mathrm{\Psi }_5(\tau ,\theta ,\overline{\theta })=\psi _5(\tau )`$) and $`s_{(a)b}^{(1)}\psi _\mu =0`$ (i.e. $`\mathrm{\Psi }_\mu (\tau ,\theta ,\overline{\theta })=\psi _\mu (\tau )`$). It is worthwhile to emphasize that the above solutions are one set of the simplest solutions which are of interest to us. A more general solution (than the above) might exist. Now the stage is set for the derivation of the nilpotent (anti-)BRST transformations for the target space coordinate variable $`x_\mu (\tau )`$. One of the most important relations, that plays a pivotal role in the derivation of the mass-shell condition ($`p^2m^2=0`$) for the Lagrangian $`L_s^{(m)}`$, is $`\dot{x}_\mu (\tau )=e(\tau )p_\mu (\tau )i\chi \psi _\mu `$. This is due to the fact that $`(L_s^{(m)}/e)=0`$ implies that $`e^2m^2=(\dot{x}_\mu +i\chi \psi _\mu )^2`$ and $`p_\mu =(L_s^{(m)}/\dot{x}^\mu )=e^1(\dot{x}_\mu +i\chi \psi _\mu )`$. A simple way to derive the (anti-)BRST transformations for the coordinate target variable $`x_\mu (\tau )`$ is to require the invariance of this central relation on the supermanifold, as $$\begin{array}{ccc}& & \dot{X}_\mu (\tau ,\theta ,\overline{\theta })E(\tau ,\theta ,\overline{\theta })P_\mu (\tau ,\theta ,\overline{\theta })+iK(\tau ,\theta ,\overline{\theta })\mathrm{\Psi }_\mu (\tau ,\theta ,\overline{\theta })\hfill \\ & =& \dot{x}_\mu (\tau )e(\tau )p_\mu (\tau )+i\chi (\tau )\psi _\mu (\tau ),\hfill \end{array}$$ $`(3.16)`$ where $`E(\tau ,\theta ,\overline{\theta })`$ is the expansion in (3.6) which has been obtained after the application of the horizontality condition. Exploiting the relations $`P_\mu (\tau ,\theta ,\overline{\theta })=p_\mu (\tau )`$ from (3.10) and $`K(\tau ,\theta ,\overline{\theta })=\chi (\tau )`$, it can be seen that the following relations emerge from (3.16): $$\begin{array}{ccc}\dot{\overline{R}_\mu }=\dot{\overline{c}}p_\mu ,\dot{R}_\mu =\dot{c}p_\mu ,\dot{S}_\mu =\dot{b}p_\mu .\hfill & & \end{array}$$ $`(3.17)`$ At this crucial stage, we summon one of the most decisive physical insights into the characteristic features of a free spinning relativistic particle which states that there is no action of any kind of force (i.e. $`\dot{p}_\mu (\tau )=0`$) on the free motion of the particle. Having taken into account this decisive input, we obtain from (3.17), the following relations $$\begin{array}{ccc}\dot{\overline{R}}_\mu _\tau \overline{R}_\mu =_\tau (\overline{c}p_\mu ),\dot{R}_\mu _\tau R_\mu =_\tau (cp_\mu ),\dot{S}_\mu _\tau S_\mu =_\tau (bp_\mu ),\hfill & & \end{array}$$ $`(3.18)`$ which lead to $$\begin{array}{ccc}\overline{R}_\mu (\tau )=\overline{c}p_\mu ,R_\mu (\tau )=cp_\mu ,S_\mu (\tau )=bp_\mu .\hfill & & \end{array}$$ $`(3.19)`$ The insertions of these values into the expansion (3.8) lead to the derivation of the nilpotent (anti-)BRST transformations ($`s_{(a)b}^{(1)}`$) on the target space coordinate field $`x_\mu (\tau )`$, as $$\begin{array}{ccc}X_\mu (\tau ,\theta ,\overline{\theta })\hfill & =& x_\mu (\tau )+\theta (s_{ab}^{(1)}x_\mu (\tau ))+\overline{\theta }(s_b^{(1)}x_\mu (\tau ))+\theta \overline{\theta }(s_b^{(1)}s_{ab}^{(1)}x_\mu (\tau )).\hfill \end{array}$$ $`(3.20)`$ In our recent papers \[24-26\] on interacting 1-form (non-)Abelian gauge theories, it has been shown that there is a beautiful consistency and complementarity between the horizontality condition and the requirement of the invariance of conserved matter (super)currents on the supermanifold. The former restriction leads to the derivation of nilpotent symmetries for the gauge- and (anti-)ghost fields. The latter restriction yields such kind of transformations for the matter fields. For the case of the free spinning relativistic particle, it can be seen that the invariance of the gauge invariant and conserved quantities on the supermanifold, leads to the derivation of the transformations for the target field variables. To corroborate this assertion, we see that the conserved and gauge invariant charge $`Q_g=\frac{1}{2}(p^2m^2)`$ is the analogue of the conserved matter current of the 1-form interacting (non-)Abelian gauge theory. Since the expansion for $`P_\mu (x.\theta ,\overline{\theta })`$ is trivial (cf. (3.10)), we have to re-express the mass-shell condition (i.e. $`e^2(p^2m^2)=(\dot{x}_\mu +i\chi \psi _\mu )^2e^2m^2`$) in the language of the superfields (3.6) and $`\mathrm{\Psi }_\mu (\tau ,\theta ,\overline{\theta })=\psi _\mu (\tau ),K(\tau ,\theta ,\overline{\theta })=\chi (\tau )`$. Thus, the invariance of the conserved (super)charges on the supermanifold is: $$\begin{array}{ccc}& & [\dot{X}_\mu (\tau ,\theta ,\overline{\theta })+iK(\tau ,\theta ,\overline{\theta })\mathrm{\Psi }_\mu (\tau ,\theta ,\overline{\theta })][\dot{X}^\mu (\tau ,\theta ,\overline{\theta })+iK(\tau ,\theta ,\overline{\theta })\mathrm{\Psi }^\mu (\tau ,\theta ,\overline{\theta })]\hfill \\ & & m^2E(\tau ,\theta ,\overline{\theta })E(\tau ,\theta ,\overline{\theta })\hfill \\ & & =[\dot{x}_\mu (\tau )+i\chi (\tau )\psi _\mu (\tau )][\dot{x}^\mu (\tau )+i\chi (\tau )\psi ^\mu (\tau )]m^2e^2.\hfill \end{array}$$ $`(3.21)`$ The equality of the appropriate terms from the l.h.s. and r.h.s. leads to $$\begin{array}{ccc}& & (\dot{x}_\mu +i\chi \psi _\mu )\dot{\overline{R}^\mu }=m^2e\dot{\overline{c}},(\dot{x}_\mu +i\chi \psi _\mu )\dot{R}^\mu =m^2e\dot{c},\hfill \\ & & (\dot{x}_\mu +i\chi \psi _\mu )\dot{S}^\mu =m^2e\dot{b},\dot{R}_\mu \dot{\overline{R}^\mu }=m^2\dot{c}\dot{\overline{c}}.\hfill \end{array}$$ $`(3.22)`$ Taking the help of the key relation $`\dot{x}_\mu +i\chi \psi _\mu =ep_\mu `$, we obtain the expressions for $`\dot{R}_\mu ,\dot{\overline{R}}_\mu ,\dot{S}_\mu `$ exactly same as the ones given in (3.17) for the mass-shell condition $`p^2m^2=0`$ to be valid. Exploiting the no force (i.e. $`\dot{p}_\mu =0`$) criterion on the free motion of a spinning relativistic particle, we obtain the expressions for $`R_\mu ,\overline{R}_\mu ,S_\mu `$ in exactly the same form as given in (3.19). The insertion of these values in (3.8) leads to the same expansion as given in (3.20). This provides the geometrical interpretation for the (anti-)BRST charges as the translational generators. It should be noted that the restrictions in (3.16) and (3.21) are intertwined. However, the latter is more physical because it states the invariance of the conserved and gauge invariant mass-shell condition explicitly. As a side remark, we would like to comment on the other conserved quantity $`p\psi m\psi _5=0`$. It is straightforward to check that this conserved quantity is automatically satisfied on the supermanifold due to the fact that $`\mathrm{\Psi }_5(\tau ,\theta ,\overline{\theta })=\psi _5(\tau ),\mathrm{\Psi }_\mu (\tau ,\theta ,\overline{\theta })=\psi _\mu (\tau ),P_\mu (\tau ,\theta ,\overline{\theta })=p_\mu (\tau )`$. In fact, the results $`\mathrm{\Psi }_5(\tau ,\theta ,\overline{\theta })=\psi _5(\tau ),\mathrm{\Psi }_\mu (\tau ,\theta ,\overline{\theta })=\psi _\mu (\tau )`$ can be obtained from this very conserved quantity if we follow the same trick as we have exploited, in the above, for the conserved and gauge invariant quantity $`p^2m^2=0`$ for the derivation of $`s_{(a)b}^{(1)}`$ for $`x_\mu (\tau )`$. 4 Supergauge BRST Symmetries: Augmented Superfield Formalism We derive here the nilpotent ($`(s_{(a)b}^{(2)})^2=0`$) (anti-)BRST symmetry transformations $`s_{(a)b}^{(2)}`$ (cf. (2.13)), corresponding to the supergauge transformations in (2.4), in the framework of augmented superfield formalism. The crucial assumption here is the condition $`\beta ^2=0`$ which can be satisfied if and only if the bosonic (anti-)ghost fields $`(\overline{\beta })\beta `$ were made up of two fermionic ghost fields. In contrast to the super 1-form bosonic connection $`\stackrel{~}{V}`$, quoted in (3.3), we define here a fermionic super 1-form connection $`\stackrel{~}{}`$ as follows: $$\begin{array}{ccc}\stackrel{~}{}=dZ(\stackrel{~}{F})=d\tau K(\tau ,\theta ,\overline{\theta })+id\theta \overline{}(\tau ,\theta ,\overline{\theta })+id\overline{\theta }(\tau ,\theta ,\overline{\theta }),\hfill & & \end{array}$$ $`(4.1)`$ where the supermultiplet $`\stackrel{~}{F}(\tau ,\theta ,\overline{\theta })=(K(\tau ,\theta ,\overline{\theta }),i(\tau ,\theta ,\overline{\theta }),i\overline{}(\tau ,\theta ,\overline{\theta })`$) has three superfields and the fermionic super gauge field $`K(\tau ,\theta ,\overline{\theta })`$ has an expansion as given in (3.12). The bosonic superfields $`(\tau ,\theta ,\overline{\theta })`$ and $`\overline{}(\tau ,\theta ,\overline{\theta })`$ have the following expansions $$\begin{array}{ccc}(\tau ,\theta ,\overline{\theta })\hfill & =& \beta (\tau )+i\theta \overline{f}_2(\tau )+i\overline{\theta }f_3(\tau )+i\theta \overline{\theta }b_2(\tau ),\hfill \\ \overline{}(\tau ,\theta ,\overline{\theta })\hfill & =& \overline{\beta }(\tau )+i\theta \overline{f}_3(\tau )+i\overline{\theta }f_2(\tau )+i\theta \overline{\theta }\overline{b}_2(\tau ),\hfill \end{array}$$ $`(4.2)`$ which yield the bosonic (anti-)ghost fields $`(\overline{\beta })\beta `$ in the limit $`(\theta ,\overline{\theta })0`$ and the bosonic ($`\beta ,\overline{\beta },b_2,\overline{b}_2`$) and fermionic ($`f_2,\overline{f}_2,f_3,\overline{f}_3`$) degrees of freedom do match. The requirement of the horizontality condition (with the super exterior derivative $`\stackrel{~}{d}`$ defined in (3.3)): $$\begin{array}{ccc}\stackrel{~}{d}\stackrel{~}{}=d=0,d=d\tau _\tau ,=d\tau \chi (\tau ),d^2=0,\hfill & & \end{array}$$ $`(4.3)`$ leads to the derivation of the secondary fields in terms of the basic fields as well as the auxiliary fields. In fact, the explicit expression for $`\stackrel{~}{d}\stackrel{~}{}`$ is: $$\begin{array}{ccc}\stackrel{~}{d}\stackrel{~}{}\hfill & =& (d\tau d\theta )(i_\tau \overline{}_\theta K)i(d\theta d\theta )(_\theta \overline{})+(d\tau d\overline{\theta })(i_\tau _{\overline{\theta }}K)\hfill \\ & & i(d\theta d\overline{\theta })(_\theta +_{\overline{\theta }}\overline{})i(d\overline{\theta }d\overline{\theta })(_{\overline{\theta }}).\hfill \end{array}$$ $`(4.4)`$ The application of the horizontality condition $`\stackrel{~}{d}\stackrel{~}{}=d=0`$, leads to the following relations: $$\begin{array}{ccc}& & _\theta \overline{}=0,_{\overline{\theta }}=0,_{\overline{\theta }}\overline{}+_\theta =0,\hfill \\ & & _\theta K=i_\tau \overline{},_{\overline{\theta }}K=i_\tau .\hfill \end{array}$$ $`(4.5)`$ The first two relations, in the above, produce $`b_2=\overline{b}_2=0,f_3=\overline{f}_3=0`$. The third one leads to $`f_2+\overline{f}_2=0`$. Choosing $`f_2=\gamma `$ implies that $`\overline{f}_2=\gamma `$ and the bosonic superfields $``$ and $`\overline{}`$ become chiral- and anti-chiral superfields, respectively, with the following expansions: $$\begin{array}{ccc}& & (\tau ,\theta )=\beta (\tau )i\theta \gamma \beta (\tau )\theta (s_{ab}^{(2)}\beta (\tau )),\hfill \\ & & \overline{}(\tau ,\overline{\theta })=\overline{\beta }(\tau )+i\theta \gamma \overline{\beta }(\tau )+\overline{\theta }(s_b^{(2)}\overline{\beta }(\tau )).\hfill \end{array}$$ $`(4.6)`$ It will be noted that the Nakanishi-Lautrup fermionic ($`\gamma ^2=0`$) auxiliary field $`\gamma (\tau )`$ is not a basic dynamical field variable of the theory (cf. (2.14)). The above equations establish that $`s_b^{(2)}\beta (\tau )=0`$ and $`s_{ab}^{(2)}\overline{\beta }(\tau )=0`$. Exploiting the expressions for $`(\tau ,\theta )`$ and $`\overline{}(\tau ,\overline{\theta })`$ (cf. (4.6)) in the last two relations of (4.5), we obtain the following values for the component local secondary fields of the expansion for $`K(\tau ,\theta ,\overline{\theta })`$ in (3.12): $$\begin{array}{ccc}b_1(\tau )=i\dot{\beta }(\tau ),\overline{b}_1(\tau )=i\dot{\overline{\beta }}(\tau ),f_1(\tau )=\dot{\gamma }(\tau ).\hfill & & \end{array}$$ $`(4.7)`$ The insertions of the above values in the expansion of $`K(\tau ,\theta ,\overline{\theta })`$ yields: $$\begin{array}{ccc}K(\tau ,\theta ,\overline{\theta })\hfill & =& \chi (\tau )+\theta (s_{ab}^{(2)}\chi (\tau ))+\overline{\theta }(s_b^{(2)}\chi (\tau ))+\theta \overline{\theta }(s_b^{(2)}s_{ab}^{(2)}\chi (\tau )).\hfill \end{array}$$ $`(4.8)`$ It will be noted that the expressions in (4.6) for the bosonic superfields can also be written in an exactly the same form as (4.8) as $`s_b^{(2)}\beta (\tau )=0`$ and $`s_{ab}^{(2)}\overline{\beta }(\tau )=0`$. There is only one caveat, however. This has to do with the $`()`$ sign in the expansion of $`B(\tau ,\theta )`$ <sup>‡‡</sup><sup>‡‡</sup>‡‡ It will be noted that the explicit form of the anti-BRST transformations $`s_{ab}^{(2)}`$ for the system are: $`s_{ab}^{(2)}x_\mu =\overline{\beta }\psi _\mu ,s_{ab}^{(2)}p_\mu =0,s_{ab}^{(2)}\overline{\beta }=0,s_{ab}^{(2)}\beta =+i\gamma ,s_{ab}^{(2)}\gamma =0,s_{ab}^{(2)}\psi _\mu =i\overline{\beta }p_\mu ,s_{ab}^{(2)}\chi =i\dot{\overline{\beta }},s_{ab}^{(2)}e=2\overline{\beta }\chi `$. The key point that should be emphasized is the fact that, the bosonic nature of the (anti-)ghost fields $`(\overline{\beta })\beta `$ does not allow a change of sign between $`s_{ab}^{(2)}\beta (=+i\gamma )`$ and $`s_b^{(2)}\overline{\beta }(=+i\gamma )`$ for the symmetry of the Lagrangian to be maintained. This situation is totally opposite to the case of $`s_{(a)b}^{(1)}`$ (cf. (2.11)) where the (anti-)ghost fields are fermionic in nature, and, that is why, the change of sign in $`s_{ab}^{(1)}c=ib,s_b^{(1)}\overline{c}=ib`$ is required.. We shall dwell on it, in detail, in the conclusions part (i.e. section 5) of our present paper. It is clear that this exercise provides the geometrical interpretation for the (anti-)BRST charges $`Q_{(a)b}^{(2)}`$ as the generators (cf. (2.15)) of translations (i.e. $`\text{Lim}_{\overline{\theta }0}(/\theta ),\text{Lim}_{\theta 0}(/\overline{\theta })`$) along the Grassmannian directions $`(\theta )\overline{\theta }`$ of the three $`(1+2)`$-dimensional supermanifold. Let us focus on the derivation of the nilpotent transformations $`s_{(a)b}^{(2)}`$ for the target field variables ($`x_\mu (\tau ),p_\mu (\tau ),\psi _\mu (\tau )`$) and the Lorentz scalar fermionic field $`\psi _5(\tau )`$ in the framework of augmented superfield formalism. Here, once again, the interplay of the horizontality condition and the invariance of the conserved quantities on the supermanifold do play a very important and decisive roles. To see it clearly, let us first concentrate on the invariance of the conserved quantities (given in (3.14)) on the supermanifold. The explicit substitutions of super expansions yield the following relationships: $$\begin{array}{ccc}& & (\dot{\psi }_5+\theta \dot{\overline{B}}_5+\overline{\theta }\dot{B}_5+\theta \overline{\theta }\dot{f}_5)(\chi +i\theta \dot{\overline{\beta }}+i\overline{\theta }\dot{\beta }\theta \overline{\theta }\dot{\gamma })m=\dot{\psi }_5\chi m,\hfill \\ & & (\dot{\psi }_\mu +\theta \dot{\overline{b}_\mu }+\overline{\theta }\dot{b}\mu +\theta \overline{\theta }\dot{f}_\mu )(\chi +i\theta \dot{\overline{\beta }}+i\overline{\theta }\dot{\beta }\theta \overline{\theta }\dot{\gamma })p_\mu =\dot{\psi }_\mu \chi p_\mu ,\hfill \end{array}$$ $`(4.9)`$ where we have exploited the expansions of $`\mathrm{\Psi }_\mu (\tau ,\theta ,\overline{\theta })`$ and $`\mathrm{\Psi }_5(\tau ,\theta ,\overline{\theta })`$ given in (3.12) and have used the expansion of $`K(\tau ,\theta ,\overline{\theta })`$ from (4.8) that has been obtained after the application of the horizontality condition. It is evident that the following relations emerge between the secondary component fields and basic fields $$\begin{array}{ccc}& & \dot{B}_5=i\dot{\beta }m_\tau (i\beta m),\dot{\overline{B}_5}=i\dot{\overline{\beta }}m_\tau (i\overline{\beta }m),\hfill \\ & & \dot{f}_5=\dot{\gamma }m_\tau (\gamma m),\dot{\overline{b}_\mu }=i\dot{\overline{\beta }}p_\mu _\tau (i\overline{\beta }p_\mu ),\hfill \\ & & \dot{b}_\mu =i\dot{\beta }p_\mu _\tau (i\beta p_\mu ),\dot{f}_\mu =\dot{\gamma }p_\mu _\tau (\gamma p_\mu ),\hfill \end{array}$$ $`(4.10)`$ where, in the latter set of entries, we have used the requirement of the free motion ($`\dot{p}_\mu =0`$) of a free spinning relativistic particle. Ultimately, the insertions of the above values in the expansions (3.12), yields the following expansions in terms of $`s_{(a)b}^{(2)}`$ (cf. (2.13)): $$\begin{array}{ccc}\mathrm{\Psi }_5(\tau ,\theta ,\overline{\theta })\hfill & =& \psi _5(\tau )+\theta (s_{ab}^{(2)}\psi _5(\tau ))+\overline{\theta }(s_b^{(2)}\psi _5(\tau ))+\theta \overline{\theta }(s_b^{(2)}s_{ab}^{(2)}\psi _5(\tau )),\hfill \\ \mathrm{\Psi }_\mu (\tau ,\theta ,\overline{\theta })\hfill & =& \psi _\mu (\tau )+\theta (s_{ab}^{(2)}\psi _\mu (\tau ))+\overline{\theta }(s_b^{(2)}\psi _\mu (\tau ))+\theta \overline{\theta }(s_b^{(2)}s_{ab}^{(2)}\psi _\mu (\tau )).\hfill \end{array}$$ $`(4.11)`$ The above expansions produce the same geometrical interpretations for the symmetries $`s_{(a)b}^{(2)}`$ and the generators $`Q_{(a)b}^{(2)}`$ (i.e. the translational generators along the Grassmannian directions) as such conclusions drawn for the expansion in (4.8) for $`K(\tau ,\theta ,\overline{\theta })`$. Having obtained the super-expansions of superfields (i.e. $`K,,\overline{},\mathrm{\Psi },\mathrm{\Psi }_5`$) in terms of the local $`\tau `$-dependent ordinary basic fields in (4.6), (4.8) and (4.11), the stage is now set for the derivation of the nilpotent symmetry transformations for the einbein field $`e(\tau )`$ and the canonically conjugate target space field variables $`x_\mu (\tau )`$ and $`p_\mu (\tau )`$. It is clear that $`\dot{p}_\mu =0`$ and the mass-shell condition $`p^2m^2=0`$ are (i) supergauge invariant (i. e. $`\delta _{sg}p_\mu =0`$), and (ii) conserved quantities. Thus, their invariance on the supermanifold, once again, leads to the same conclusions as illustrated in (3.9) and (3.10). As a consequence, we have $`s_{(a)b}^{(2)}p_\mu =0`$. Now, the central problem is to obtain the nilpotent transformations for $`e(\tau )`$ and $`x_\mu (\tau )`$. In this connection, it turns out that (i) the intertwined relations given in (3.16) and (3.21), and (ii) the conserved quantity $`p\psi m\psi _5=0`$ (expressed in terms of $`p_\mu =e^1(\dot{x}_\mu +i\chi \psi _\mu )`$ so that it becomes $`\dot{x}\psi me\psi _5=0`$), emerge to help in the final computation. Exploiting the explicit expressions for the expansions of $`\mathrm{\Psi }_\mu (\tau ,\theta ,\overline{\theta })`$ and $`\mathrm{\Psi }_5(\tau ,\theta ,\overline{\theta })`$ given in (4.11) and inserting the values of $`\dot{X}_\mu (\tau ,\theta ,\overline{\theta })`$ and $`E(\tau ,\theta ,\overline{\theta })`$ from (3.8) and (3.1) in the following invariance of the conserved quantity on the supermanifold: $$\begin{array}{ccc}\dot{X}\mathrm{\Psi }mE\mathrm{\Psi }_5=\dot{x}\psi me\psi _5,\hfill & & \end{array}$$ $`(4.12)`$ we obtain the following relationships $$\begin{array}{ccc}& & \dot{\overline{R}}\psi m\overline{f}\psi _5=ie\overline{\beta }m^2i\overline{\beta }(\dot{x}p),\hfill \\ & & \dot{R}\psi mf\psi _5=ie\beta m^2i\beta (\dot{x}p),\hfill \\ & & i(\dot{S}\psi )+i\overline{\beta }(\dot{R}p)i\beta (\dot{\overline{R}}p)+i\overline{f}\beta m^2if\overline{\beta }m^2iB\psi _5\hfill \\ & & =\gamma (\dot{x}p)e\gamma m^2.\hfill \end{array}$$ $`(4.13)`$ Similarly, tapping the potential of the cute relationship in (3.16) (which is finally intertwined with the invariance of the conserved quantity in (3.21)), we obtain the following connections among the component secondary local fields and the basic local fields: $$\begin{array}{ccc}& & \dot{R}_\mu fp_\mu =\dot{\beta }\psi _\mu \beta \chi p_\mu ,\hfill \\ & & \dot{\overline{R}}_\mu \overline{f}p_\mu =\dot{\overline{\beta }}\psi _\mu \overline{\beta }\chi p_\mu ,\hfill \\ & & i\dot{S}_\mu Bp_\mu =i\dot{\gamma }\psi _\mu +i(\dot{\overline{\beta }}\beta \dot{\beta }\overline{\beta })p_\mu +i\chi \gamma p_\mu ,\hfill \end{array}$$ $`(4.14)`$ where the explicit expansions from (3.1) (for $`E`$), (3.8) (for $`X_\mu `$), (4.8) (for $`K`$) and (4.11) (for $`\mathrm{\Psi }_\mu )`$ have been used. Solution to the above equations on the on-shell (i.e. $`\dot{\psi }_\mu =\chi p_\mu ,\dot{p}_\mu =0,\dot{x}_\mu =ep_\mu i\chi \psi _\mu ,p^2m^2=0,p\psi m\psi _5=0`$, etc.) are as follows: $$\begin{array}{ccc}& & f=2\beta \chi ,\overline{f}=2\overline{\beta }\chi ,R_\mu =\beta \psi _\mu ,\overline{R}_\mu =\overline{\beta }\psi _\mu ,\hfill \\ & & S_\mu =\gamma \psi _\mu +\beta \overline{\beta }p_\mu ,B=2\dot{\beta }\overline{\beta }+2\gamma \chi .\hfill \end{array}$$ $`(4.15)`$ As a side remark, it is interesting to point out that, even from a single relationship in (4.14), some of the above values could be guessed. For instance, the relationship $`\dot{R}_\mu fp_\mu =\dot{\beta }\psi _\mu \beta \chi p_\mu `$ can be re-expressed as $`\dot{R}_\mu fp_\mu =_\tau (\beta \psi _\mu )\beta \dot{\psi }_\mu \beta \chi p_\mu `$. Exploiting the on-shell condition $`\dot{\psi }_\mu =\chi p_\mu `$ in the above, it can be seen that $`\dot{R}_\mu fp_\mu =_\tau (\beta \psi _\mu )2\beta \chi p_\mu `$. This last relation gives a glimpse of $`R_\mu =\beta \psi _\mu `$ and $`f=2\beta \chi `$. In exactly the same manner, it can be seen that $`\overline{R}_\mu =\overline{\beta }\psi _\mu `$ and $`\overline{f}=2\overline{\beta }\chi `$. With these values, other expressions of (4.15) follow, which ultimately, satisfy all the relations derived in (4.13) and (4.14). Insertions of the above values in the expressions for superfield $`E(\tau ,\theta ,\overline{\theta })`$ in (3.1) and the superfield $`X_\mu (\tau ,\theta ,\overline{\theta })`$ in (3.8) lead to the following: $$\begin{array}{ccc}E(\tau ,\theta ,\overline{\theta })\hfill & =& e(\tau )+\theta (s_{ab}^{(2)}e(\tau ))+\overline{\theta }(s_b^{(2)}e(\tau ))+\theta \overline{\theta }(s_b^{(2)}s_{ab}^{(2)}e(\tau )),\hfill \\ X_\mu (\tau ,\theta ,\overline{\theta })\hfill & =& x_\mu (\tau )+\theta (s_{ab}^{(2)}x_\mu (\tau ))+\overline{\theta }(s_b^{(2)}x_\mu (\tau ))+\theta \overline{\theta }(s_b^{(2)}s_{ab}^{(2)}x_\mu (\tau )).\hfill \end{array}$$ $`(4.16)`$ The above equation, once again, establishes the fact that the nilpotent (anti-)BRST generators $`Q_{(a)b}^{(2)}`$ (and corresponding symmetry transformations $`s_{(a)b}^{(2)}`$) are the translational generators along the Grassmannian directions $`(\theta )\overline{\theta }`$ of the supermanifolds. 5 Conclusions In our present endeavour, we have exploited, in an elegant way, the key ideas of the augmented superfield formalism to derive two sets of anticommuting (i.e. $`\{s_{(a)b}^{(1)},s_{(a)b}^{(2)}\}=0`$) and nilpotent ($`(s_{(a)b}^{(1,2)})^2=0`$) (anti-)BRST symmetry transformations $`s_{(a)b}^{(1,2)}`$ for all the fields variables, present in the Lagrangian description of a free massive spinning relativistic particle. The theoretical arsenal of (i) the horizontality condition, and (ii) the invariance of the conserved quantities on the supermanifold, have played very decisive roles in the above derivations. One of the central new features of our present investigation is the application of the augmented superfield formulation to a supersymmetric system where the bosonic $`(\overline{\beta })\beta `$ and fermionic $`(\overline{c})c`$ (anti-)ghost fields are present together in the (anti-)BRST invariant Lagrangian describing the free motion ($`\dot{p}_\mu =0`$) of the super-particle. Of course, for the system under consideration, the reparametrization symmetry invariance and the gauge symmetry invariance are also present. All these symmetries are inter-related. Even though the pair of bosonic (anti-)ghost fields $`(\overline{\beta })\beta `$ are commutative in nature (i.e. $`\beta \overline{\beta }=\overline{\beta }\beta ,\beta \mathrm{\Sigma }=\mathrm{\Sigma }\beta ,\overline{\beta }\mathrm{\Sigma }=\mathrm{\Sigma }\overline{\beta }`$ for the generic field $`\mathrm{\Sigma }=x_\mu ,p_\mu ,\psi _\mu ,\psi _5,e,\chi ,c,\overline{c},b,\gamma `$), they are taken to be nilpotent of order two ($`\beta ^2=0,\overline{\beta }^2=0`$) with the assumption that they are made up of a pair of fermionic (anti-)ghost fields (i.e. $`\beta c_1c_2,\overline{\beta }\overline{c}_1\overline{c}_2,c_1^2=c_2^2=0,c_1c_2+c_2c_1=0`$, etc.). Such an assumption is essential for a couple of advantageous reasons. First, the (anti-)BRST transformations (3.13) (corresponding to the supergauge symmetry transformations (2.4)) become nilpotent (i.e. $`(s_{(a)b}^{(2)})^2=0`$) under the above assumption. This can be explicitly checked for $`(s_{(a)b}^{(2)})^2x_\mu (\tau )=0`$ and $`(s_{(a)b}^{(2)})^2e(\tau )=0`$ where the conditions $`\beta ^2=\overline{\beta }^2=0`$ and $`_\tau (\beta )^2=_\tau (\overline{\beta })^2=0`$ are required for the proof of an explicit nilpotency. Second, this assumption also allows $`s_{(a)b}^{(2)}`$ to decouple from the nilpotent transformations in (2.6) (where, in some sense, they are hidden) and the nilpotent transformations (2.11) so that they could become completely separate and independent. At this stage, it is worth emphasizing that there is no such kind of restriction (i.e. $`\beta ^20`$) on the bosonic ghost field $`\beta `$ in the nilpotent transformations listed in (2.6). In our earlier works \[23-27\], the horizontality condition was augmented to include the invariance of the conserved matter currents/charges on the supermanifold. In our present endeavour, the augmented superfield formalism has been extended to include the invariance of any kind of conserved quantities on the supermanifold and still (i) there is a mutual consistency and complementarity between the two above types of restrictions. (ii) The geometrical interpretations for the nilpotent (anti-)BRST charges $`Q_{(a)b}^{(1,2)}`$, as the translational generators $`(\text{Lim}_{\overline{\theta }0}(/\theta ))\text{Lim}_{\theta 0}(/\overline{\theta })`$ along the $`(\theta )\overline{\theta }`$-directions of the $`(D+2)`$-dimensional supermanifold, remains intact. (iii) The nilpotency of the (anti-)BRST charges $`Q_{(a)b}^{(1,2)}`$ is encoded in a couple of successive translations (i.e. $`(/\theta )^2=(/\overline{\theta })^2=0`$) along either of the two Grassmannian directions of the supermanifold. (iv) The anticommutativity of the nilpotent (anti-)BRST charges $`Q_{(a)b}^{(1,2)}`$ (and the transformations they generate) is captured in the relationship $`(/\theta )(/\overline{\theta })+(/\overline{\theta })(/\theta )=0`$. Thus, our present extension of our earlier works \[23-27\] is a very natural generalization of the horizontality condition where the beauty of the geometrical interpretations is not spoiled in any way. We dwell a bit on the negative sign in (4.6) for the expansion of the chiral superfield $`(\tau ,\theta )`$. In fact, the bosonic ghost term $`\dot{\overline{\beta }}\dot{\beta }`$ in (2.7) (or (2.14)) remains invariant under $`\beta \pm \overline{\beta },\overline{\beta }\pm \beta `$. We have taken the $`(+)`$ sign for our description of the anti-BRST transformations $`s_{ab}^{(2)}`$. However, one could choose $`\beta \overline{\beta },\overline{\beta }\beta `$ equally well. In that case, the negative sign in the expansion of (4.6) disappears. However, under the latter choice, the beautiful expansions of the superfields $`E(\tau ,\theta ,\overline{\theta })`$ and $`X_\mu (\tau ,\theta ,\overline{\theta })`$ in (4.16) get disturbed and some minus signs crop up in the expansion. This is why, we have opted for the $`(+)`$ sign in ($`\beta \pm \overline{\beta },\overline{\beta }\pm \beta `$) and all the transformations, listed in the whole body of the present text, are consistent with it. Geometrically, it seems that the translation of the anti-chiral superfield $`\overline{}(\tau ,\overline{\theta })`$ along the $`(+\overline{\theta })`$ direction of the supermanifold produces the BRST $`s_b^{(2)}`$ transformation for the anti-ghost field $`\overline{\beta }`$. However, the anti-BRST transformation $`s_{ab}^{(2)}`$ for the ghost field $`\beta `$ is produced by the translation of the chiral superfield $`(\tau ,\theta )`$ along the $`(\theta )`$ direction of the supermanifold. This kind of discrepancy appears, perhaps, because of the peculiar behaviour of these bosonic (anti-)ghost fields which are commutative in nature but are restricted to be nilpotent of order two (i.e. $`\beta ^2=0,\overline{\beta }^2=0`$). In our present investigation, we have not dwelt on the derivation of the beautiful nilpotent symmetries $`s_{(a)b}^{(0)}`$ (cf. (2.6)) in the framework of augmented superfield formalism because we cannot add the bosonic and fermionic 1-form super connections $`\stackrel{~}{V}`$ and $`\stackrel{~}{}`$ (defined in (3.3) and (4.1)) together. However, we strongly believe that if (i) the action with Lagrangian (2.7) is written in terms of the superfields, and (ii) the BRST symmetry transformations (2.6) are expressed in terms of the superfield too, we shall be able to derive these beautiful nilpotent symmetry transformations (i.e. without any restrictions) in the framework of augmented superfield approach to BRST formalism. We would like to lay stress on the fact that the nilpotent symmetry transformations in (2.6) are beautiful because there are no restrictions (i.e. $`\beta ^20,\overline{\beta }^20`$) and no other peculiarities (like its being a composite of two fermions, etc.) are associated with the bosonic (anti-)ghost fields $`(\overline{\beta })\beta `$. Furthermore, our approach could be extended to be applied to some more complicated and interesting field theoretic supersymmetric systems so that, the ideas proposed in our present endeavour, could be put on a firmer footing. These are some of the issues that are under investigation and we shall report these results in our forthcoming publications . Acknowledgements The warm hospitality extended by the HEP group of the AS-ICTP, Trieste is gratefully acknowledged. It is a great pleasure to thank L. Bonora (SISSA, Trieste), E. Ivanov (JINR, Dubna), K. S. Narain (AS-ICTP) and G. Thompson (AS-ICTP) for fruitful discussions.
warning/0506/cond-mat0506341.html
ar5iv
text
# Self-avoiding walks crossing a square ## 1 Introduction We consider the problem of self-avoiding walks on the square lattice $`^2.`$ For walks on an infinite lattice, it is generally accepted that the number of such walks of length $`n`$, equivalent up to a translation, denoted $`c_n`$, grows as $`c_nconst.\mu ^nn^{\gamma 1},`$ with metric properties, such as mean-square radius of gyration or mean-square end-to-end distance growing as $`R^2_nconst.n^{2\nu },`$ where $`\gamma =43/32`$ and $`\nu =3/4.`$ The growth constant $`\mu `$ is lattice dependent, and for the square lattice is not known exactly, but is indistinguishable numerically from the unique positive root of the equation $`13x^47x^2581=0.`$ We denote the generating function by $`C(x):=_nc_nx^n.`$ It will be useful to define a second generating function for those SAW which start at the origin $`(0,0)`$ and end at a given point $`(u,v),`$ as $`G_{(0,0;u,v)}(x).`$ In terms of this generating function, the mass $`m(x)`$ is defined to be the rate of decay of $`G`$ along a coordinate axis, $$m(x):=\underset{n\mathrm{}}{lim}\frac{\mathrm{log}G_{(0,0;n,0)}(x)}{n}.$$ (1) Here, we are interested in a restricted class of square lattice SAW which start at the origin $`(0,0)`$, end at $`(L,L),`$ and are entirely contained in the square $`[0,L]\times [0,L].`$ A fugacity, or weight, $`x`$ is associated with each step of the walk. Historically, this problem seems to have led two largely independent lives. One as a problem in combinatorics (in which case the fugacity has been implicitly set to $`x=1`$), and one in the statistical mechanics literature where the behaviour as a function of fugacity $`x`$ has been of considerable interest, as there is a fugacity dependent phase transition. The problem seems to have first been seriously studied as a mathematical problem by Abbott and Hanson in 1978, many of whose results and methods are still powerful today. A key question considered both then and now, is the number of distinct SAW on the constrained lattice, and their growth as a function of the size of the lattice. Let $`c_n(L)`$ denote the number of $`n`$-step SAW which start at the origin $`(0,0)`$, end at $`(L,L)`$ and are entirely contained in the square $`[0,L]\times [0,L].`$ Further, let $`C_L(x):=_nc_n(L)x^n.`$ Then $`C_L(1)`$ is the number of distinct walks from the origin to the diagonally opposite corner of an $`L\times L`$ lattice. In , and independently in , it was proved that $`C_L(1)^{1/L^2}\lambda .`$ The value of $`\lambda `$ is not known, though bounds and estimates have been given in . One of our purposes in this paper is to improve on both the bounds and the estimate. Like so many problems in lattice statistics, this one owes a debt to J. M. Hammersley. A closely related problem to the one considered here is discussed in , which is in turn devoted to problems posed by Hammersley. However the earliest mention of this problem appears to be by Knuth , who calculated the number of SAW crossing a $`10\times 10`$ square by Monte Carlo methods, and estimated the number to be $`(1.6\pm 0.3)\times 10^{24}.`$ It is now known, see Table 2 below, that the correct answer is $`1.5687..\times 10^{24}.`$ A related problem was studied by Edwards in . He considered SAW starting at a point denoted the origin with end point a distance $`L`$ from the origin, and no other points at distance $`L`$ or greater. Let $`g(L)`$ denote the number of such SAW. Then Edwards proved that $`lim_L\mathrm{}g(L)^{(1/L^2)}`$ exists and lies between 2.3 and 5.0. In our notation, Edwards has proved that $`1.53<\lambda <2.24.`$ Edwards also proved that the same limit holds for SAW from the origin to the boundary of any convex, bounded subset of $`^2.`$ His numerical work led him to suggest that $`\lambda `$ is about 1.77. Our best estimate, given below, is 1.744550(5). The problem of Hamiltonian paths on an $`L\times M`$ rectangular grid, going from $`(0,0)`$ to $`(L,M)`$ has also been considered previously. Earlier work is described in , where Collins and Krompart also give generating functions for the number of such paths on grids with $`M=1,2,3,4,5.`$ In Jacobsen and Kondev gave a field-theoretical estimate of the growth constant for Hamiltonian SAW on the square lattice, which must fill a square, as $`1.472801\pm 0.00001`$. In the statistical mechanics literature, the problem appears to have been introduced by Whittington and Guttmann in 1990, who were particularly interested in the phase transition that takes place as one varies the fugacity associated with the walk length. All walks on lattices up to $`6\times 6`$ were enumerated, and the estimate $`\lambda =1.756\pm 0.01`$ was given. At a critical value, $`x_c`$ the average walk length of a path on an $`L\times L`$ lattice changes from $`\mathrm{\Theta }(L)`$ to $`\mathrm{\Theta }(L^2),`$ where we define $`\mathrm{\Theta }(x)`$ as follows: Let $`a(x)`$ and $`b(x)`$ be two functions of some variable $`x`$. We write that $`a(x)=\mathrm{\Theta }(b(x))`$ as $`xx_0`$ if there exist two positive constants $`\kappa _1`$ and $`\kappa _2`$ such that, for $`x`$ sufficiently close to $`x_0`$, $$\kappa _1b(x)a(x)\kappa _2b(x).$$ In the critical fugacity was proved to be at least $`1/\mu `$, its value was estimated numerically and was conjectured to be $`x_c=1/\mu `$, and in the conjecture was proved by Madras. The problem was subsequently taken up by Burkhardt and Guim , who extended the enumerations given in to $`9\times 9`$ lattices, and used their data to give the improved estimate $`\lambda =1.743\pm 0.005.`$ By considering SAW as the $`N0`$ limit of the O$`(N)`$ model of magnetism, Burkhardt and Guim show that the conjecture $`x_c=1/\mu `$ made in on numerical grounds follows directly, though this is not a proof, unlike the subsequent result of Madras . They also gave a scaling Ansatz for the behaviour of $`C_L(x)`$ for $`L`$ large in the vicinity of $`x=x_c.`$ They proposed $$C_L(x)L^{\eta _c}f[L^{1/\nu }(x_cx)]$$ (2) where $`\nu =3/4`$, as described above, and $`\eta _c=5/2`$ is the corner exponent of the magnetisation , given by Cardy’s result $`\eta _c(\theta )=\frac{\pi }{\theta }\eta _{},`$ for a wedge-angle $`\theta `$, which is $`\pi /2`$ in this case. $`\eta _{}=5/4`$ is the surface exponent that characterises the decay of spin-spin correlations parallel to the boundary in the semi-infinite geometry, corresponding to wedge-angle $`\pi .`$ Consequences of this scaling Ansatz include the following predictions: $$C_L(x_c)const.L^{\eta _c}$$ $$n(x_c,L)=x\frac{}{x}C_L(x_c)const.L^{1/\nu }$$ (3) $$(n(x_c,L)n(x_c,L))^2=(x\frac{}{x})^2\mathrm{ln}C_L(x_c)const.L^{2/\nu }.$$ They tested these results from their numerical data, and found them well supported. We provide even firmer support for these results on the basis of radically extended numerical data. Equation (3) has also previously been given by Duplantier and Saleur . Burkhardt and Guim also considered a generalisation of the problem considered here by including a second fugacity, associated with steps in the boundary. This allows the problem of adsorbing boundaries to be studied. We will not discuss this aspect of the problem further, except to note that in a full scaling theory is developed, and the predictions of the theory are tested against numerical data. In the slightly more general problem of SAW constrained to an $`L\times M`$ lattice was considered, where the analogous question was asked: how many non-self-intersecting paths are there from $`(0,0)`$ to $`(L,M)\mathrm{?}`$ If one denotes the number of such paths by $`C_{L,M},`$ it is clear that, for $`M`$ finite, the paths can be generated by a finite dimensional transfer matrix, and hence that the generating function is rational . Indeed, in it was proved that $$G_2(z)=\underset{L0}{}C_{L,2}z^L=\frac{1z^2}{14z+3z^22z^3z^4},$$ (4) (where here we have corrected a typographical error). It follows that $`C_{L,2}\mathrm{const}.\lambda _2^{2L},`$ where $`\lambda _2=\sqrt{\frac{2}{\sqrt{13}3}}=1.81735\mathrm{}`$. In this paper we also consider two further problems which can be seen as generalisations of the stated problem. Firstly, we consider the problem where SAWs are allowed to start anywhere on the left edge of the square and terminate anywhere on the right edge; so these are walks traversing the square from left to right. We call such walks transverse walks. Secondly, we consider the problem in which there may be several independent SAW, each SAW starting and ending on the perimeter of the square. The SAW are not allowed to take steps along the edges of the perimeter. Such walks partition the square into distinct regions and by colouring the regions alternately black and white we get a cow-patch pattern. Each problem is illustrated in Figure 1. Following the work in , Madras proved a number of theorems. In fact, most of Madras’s results were proved for the more general $`d`$-dimensional hyper-cubic lattice, but here we will quote them in the more restricted two-dimensional setting. ###### Theorem 1 The following limits, $$\mu _1(x):=\underset{L\mathrm{}}{lim}C_L(x)^{1/L}\text{ and }\mu _2(x):=\underset{L\mathrm{}}{lim}C_L(x)^{1/L^2},$$ are well-defined in $`\text{}\{+\mathrm{}\}`$. More precisely, * $`\mu _1(x)`$ is finite for $`0<x1/\mu `$, and is infinite for $`x>1/\mu .`$ Moreover, $`0<\mu _1(x)<1`$ for $`0<x<1/\mu `$ and $`\mu _1(1/\mu )=1.`$ * $`\mu _2(x)`$ is finite for all $`x>0.`$ Moreover, $`\mu _2(x)=1`$ for $`0<x1/\mu `$ and $`\mu _2(x)>1`$ for $`x>1/\mu .`$ In the existence of the limit $`\mu _2(x)`$ was proved, and in addition upper and lower bounds on $`\mu _2(x)`$ were established. The average length of a (weighted) walk is defined to be $$n(x,L):=\underset{n}{}nc_n(L)x^n/\underset{n}{}c_n(L)x^n.$$ (5) ###### Theorem 2 For $`0<x<1/\mu `$, we have that $`n(x,L)=\mathrm{\Theta }(L)`$ as $`L\mathrm{}`$, while for $`x>1/\mu `$, we have $`n(x,L)=\mathrm{\Theta }(L^2)`$. In it was proved that $`n(1)_L=\mathrm{\Theta }(L^2)`$. The situation at $`x=1/\mu `$ is unknown. We provide compelling numerical evidence that in fact $`n(1/\mu )_L=\mathrm{\Theta }(L^{1/\nu })`$ , where $`\nu =3/4`$, in accordance with an intuitive suggestion in both and . ###### Theorem 3 For $`x>0`$, define $`f_1(x)=\mathrm{log}\mu _1(x)`$ and $`f_2(x)=\mathrm{log}\mu _2(x).`$ * The function $`f_1`$ is a strictly increasing, negative-valued convex function of $`\mathrm{log}x`$ for $`0<x<1/\mu ,`$ and $`f_1(x)=\mathrm{\Theta }(m(x))`$ as $`x1/\mu ^{}`$, where $`m(x)`$ is the mass, defined by (1). * The function $`f_2`$ is a strictly increasing, convex function of $`\mathrm{log}x`$ for $`x>1/\mu ,`$ and satisfies $`0<f_2(x)\mathrm{log}\mu +\mathrm{log}x.`$ Some, but not all of the above results were previously proved in , but these three theorems elegantly capture all that is rigorously known. The rest of the paper is organised as follows: In the next section we describe our enumeration methods, and explain how they are used to obtain radically extended series expansions for the number of walks crossing a square, the number of cow-patch configurations and the number of transverse SAW. Section 3 details the results we have obtained. In Section 4 we derive methods for obtaining rigorous upper and lower bounds on $`\lambda .`$ In that section we show that upper bounds based on counting cow-patch configurations are fully equivalent to the method of Abbott and Hanson, based on 0–1 admissible matrices. An improved method of lower bounds based on counting transverse walks is also derived. In section 5 we then apply these methods to our radically extended enumerations to provide significantly improved bounds on $`\lambda .`$ In section 6 we give exact results for short SAW crossing a square. The shortest SAW that can cross a square from $`(0,0)`$ to $`(L,L)`$ is of length $`2L.`$ We give the exact number of such SAW of length $`2L+2K,`$ for $`K=0,1,2,`$ and asymptotic results for $`K=\mathrm{o}(L^{1/3})`$. Section 7 is devoted to a numerical analysis which gives precise (though non-rigorous) estimates of $`\lambda ,`$ for all three types of configurations, a discussion of the mean number of steps as a function of fugacity, fluctuations in this quantity, and a scaling theory for such fluctuations. We also speculate on the nature of the sub-dominant behaviour of the asymptotic form for the number of SAW. Section 8 is also a numerical study, but of the number of SAW that pass through the central vertex of an $`L\times L`$ square. Finally in section 9 we study Hamiltonian paths, obtaining both rigorous upper and lower bounds on the growth constant, and a numerical estimate. ## 2 Exact enumeration In the following we give a fairly detailed description of the algorithm we use to enumerate the number of walks crossing a square and briefly outline how this basic algorithm is modified in order to include a step fugacity, study SAWs traversing a square and the cow-patch configurations. ### 2.1 The basic algorithm We use a transfer matrix algorithm to count the number of walks crossing $`L\times M`$ rectangles. The algorithm is based on the method of Conway et al. for enumerating ordinary self-avoiding walks. The transfer matrix technique involves drawing a boundary line through the rectangle intersecting $`M+1`$ or $`M+2`$ edges. For each configuration of occupied or empty edges we maintain a count of partially completed walks intersecting the boundary in that pattern. Walks in rectangles are counted by moving the boundary adding one vertex at a time (see Figure 2). Rectangles are built up column by column with each column constructed one vertex at a time. Configurations are represented by lists of states $`\{\sigma _i\}`$, where the value of the state $`\sigma _i`$ at position $`i`$ must indicate if the edge is occupied or empty. An empty edge is indicated by $`\sigma _i=0`$. An occupied edge is either free (not connected to other edges) or connected to exactly one other edge via a path to the left of the boundary. We indicate this by $`\sigma _i=1`$ for a free end, $`\sigma _i=2`$ for the lower end of a loop and $`\sigma _i=3`$ for the upper end of a loop connecting two edges. Since we are studying self-avoiding walks on a two-dimensional lattice the compact encoding given above uniquely specifies which ends are paired. Read from the bottom the configuration along the intersection in Figure 2 is $`\{2203301203\}`$ (prior to the move) and $`\{2300001203\}`$ (after the move). There are major restrictions on the possible configurations and their updating rules. Firstly, since the walk has to cross the rectangle there is exactly one free end in any configuration. Secondly, all remaining occupied edges are connected by a path to the left of the intersection and we cannot close a loop. It is therefore clear that the total number of 2’s equals the total number of 3’s. Furthermore, as we look through the configuration from the bottom the number of 2’s is never smaller than the number of 3’s (they are perfectly balanced parentheses). We also have to ensure that the graphs we construct have only one connected component. In the following we shall briefly show how this is achieved. We call the configuration before and after the move the ‘source’ and ‘target’, respectively. Initially we have just one configuration with a single ‘1’ at position $`0`$ (all other entries ‘0’) thus ensuring that we start in the bottom-left corner. As the boundary line is moved one step, we run through all the existing sources. Each source gives rise to one or two targets and the count of the source is added to the count of the target (the initial count of a target being zero). After a source has been processed it can be discarded since it will make no further contribution. Table 1 lists the possible local ‘input’ states and the ‘output’ states which arise as the kink in the boundary is propagated one step and the various symbols will be explained below. Firstly, the values of the ’Bottom’ and ’Top’ table entries refer to the edge-states of the kink prior to the move. The Top (Bottom) entry is the state of the edge intersected by the horizontal (lower vertical) part of the boundary. Some of the updating rules are illustrated further in Figure 2. The topmost panels represent the input state ‘00’ having the allowed output states ‘00’ and ‘23’ corresponding to leaving the edges empty or inserting a new loop, respectively. The middle panels represents the input state ‘20’ with output states ‘20’ and ‘02’ from the two ways of continuing the loop end (note that the loop has to be continued since we would otherwise generate an additional free end not located at the allowed positions in the corners). The bottommost panels represents the input state ‘22’ as part of the configuration $`\{02233\}`$. In this case we connect two loop ends and we thus join two separate loops into a single larger loop. The matching upper end of the innermost loop becomes the new lower end of the joined loop. The relabeling of the matching loop-end when connecting two ‘2’s (or two ‘3’s) is denoted by over-lining in Table 1. When we join loop ends to a free end (inputs ‘12’, ‘21’, ‘13’, and ’31’) we have to relabel the matching loop end as a free end. This type of relabeling is indicated by the symbol $`\widehat{00}`$. The input state ‘11’ never occurs since there is only one free end. The input state ‘23’ is not allowed since connecting the two ends results in a closed loop and we thus discard any configuration in which a closed loop is formed. It is quite easy to avoid forming closed loops. We only have to be careful when the input is ’03’ or ’30’. If the upper end of the loop is continued along the vertical output edge we would form a closed loop if the horizontal edge immediately below was a lower loop end, and we just check the state of this edge and only proceed if it is not in state ’2’ (naturally the upper loop end can always by continued along the horizontal output edge). Finally, we have marked two outputs, from the inputs ‘01’ and ‘10’ with ‘Res’, indicating situations where we terminate free ends. This results in completed partial walks and is only allowed if there are no other occupied edges in the source (otherwise we would produce graphs with separate pieces) and if we are at the top-most vertex (otherwise we would not cross the rectangle). The count for this configuration is the number of walks crossing a rectangle of height $`M`$ and length $`L`$ equal to the number of completed columns. The time required to obtain the number of walks on $`L\times M`$ rectangles grows exponentially with $`M`$ and linearly with $`L`$. Time and memory requirements are basically proportional to the maximal number of distinct configurations along the boundary line. When there is no kink in the intersection (a column has just been completed) we can calculate this number, $`N_{\mathrm{conf}}(M)`$, exactly. Obviously the free end cuts the boundary line configuration into two separate pieces. Each of these pieces consists of ‘0’s and an equal number of ‘2’s and ‘3’s with the latter forming a perfectly balanced parenthesis system. Each piece thus corresponds to a Motzkin path \[17, Ch. 6\] (just map 0 to a horizontal step, 2 to a north-east step, and 3 to a south-east step). The number of Motzkin paths $`M_n`$ with $`n`$ steps is easily derived from the generating function $`(x)=_nM_nx^n`$, which satisfies $`=1+x+x^2^2`$, so that $$(x)=[1x\sqrt{(1+x)(13x)}]/2x^2.$$ (6) The number of configurations $`N_{\mathrm{conf}}(M)`$ for a rectangle of height $`M`$ is simply obtained by inserting a free end between two Motzkin paths, so that the generating function $`_MN_{\mathrm{conf}}(M)x^M`$ is simply $`x(x)^2`$. The Lagrange inversion formula gives $$N_{\mathrm{conf}}(M)=2\underset{i0}{}\frac{(M+1)!}{i!(i+2)!(M2i)!}.$$ When the boundary line has a kink the number of configurations exceeds $`N_{\mathrm{conf}}(M)`$ but clearly is less than $`N_{\mathrm{conf}}(M+1)`$. From (6) we see that asymptotically $`N_{\mathrm{conf}}(M)`$ grows like $`3^M`$ (up to a power of $`M`$). So the same is true for the maximal number of boundary line configurations and hence for the computational complexity of the algorithm. Note that the total number a walks grows like $`\lambda ^{LM},`$ so our algorithm leads to a better than exponential improvement over direct enumeration. The integers occurring in the expansion become very large so the calculation was performed using modular arithmetic . This involves performing the calculation modulo various prime numbers $`p_i`$ and then reconstructing the full integer coefficients at the end. We used primes of the form $`p_i=2^{30}r_i`$ where $`r_i`$ are distinct integers, less than 1000, such that $`p_i`$ is a (different) prime for each value of $`i`$. The Chinese remainder theorem ensures that any integer has a unique representation in terms of residues. If the largest integer occurring in the final expansion is $`m`$, then we have to use a number of primes $`k`$ such that $`p_1p_2\mathrm{}p_k>m`$. ### 2.2 Extensions of the algorithm The algorithm is easily generalised to include a step fugacity $`x`$. The count associated with the boundary line configuration has to be replaced by a generating function for partial walks. Since we only use this generalisation to study walks crossing an $`L\times L`$ square the generating function is just a polynomial of degree (at most) $`L(L+2)`$ in $`x`$. The coefficient of $`x^n`$ is just the number of partial walks of length $`n`$ intersecting the boundary line in the pattern specified by the configuration. The generating function of the source is multiplied by $`x^m`$ and added to the target, where $`m`$ is the number of additional steps inserted. Not all $`L(L+2)`$ terms in the polynomials need be retained. Firstly, any walk crossing the square has even length. Thus in the generating functions for partial walks either all the even or all the odd terms are zero, and we need only retain the non-zero terms. Secondly, in order to construct a given boundary line configuration, a certain minimal number of steps $`n_{\mathrm{min}}`$ are required. Terms in the generating function of degree lower than $`n_{\mathrm{min}}`$ are therefore zero and again we need not store these. The generalisation to traversing walks is also quite simple. Firstly, we have $`M+1`$ initial configurations which are empty except for a free end at position $`0jM`$. This corresponds to the $`M+1`$ possible starting positions for the walk on the left boundary. Secondly, we have to change how we produce the final counts. The easiest way to ensure that a walk spans the rectangle and that only single component graphs are counted is as follows: When column $`L+1`$ has been completed we look at the $`M+1`$ configurations with a single free end and add the counts from all of them. This is the number of walks traversing an $`L\times M`$ rectangle. The generalisation to cow-patch patterns is more complicated. Graphs can now have many separate components each of which is a SAW, and there can thus be many free ends in a boundary line configuration. Note that each SAW starts and terminates with a step perpendicular to the border of the rectangle and there are never any steps along the edges of the borders of the rectangle. There are $`2^{M1}`$ initial configurations since any of the edges in the first column from position 1 to $`M1`$ can be occupied by a free end or be empty (recall that in cow-patch configurations the top and bottom-most horizontal edges cannot be occupied). There is an extra updating rule in the bulk in that we can have the local input ‘11’ (joining of two free ends) with the only possible output being ‘00’. Also the updating rules at the upper and lower borders of the rectangle are different in this case. At the upper border we only have the input ‘00’ with the outputs ‘00’ and ‘10’ corresponding to the insertion of a free end on a vertical edge at the upper border. There is no ‘23’ or ‘01’ outputs since these would produce an occupied edge along the upper border. At the lower border we have inputs ‘00’, ‘01’, and ‘02’ and in each case the only possible output is ‘00’ (with the appropriate relabeling in the ‘02’ case). Finally, the count of the number of cow-patch patterns is obtained by summing over all boundary line configurations after the completion of a column. ## 3 Results As discussed above, in order to obtain the exact value of the number of SAW crossing a square, some of which are integers with nearly 100 digits, we performed the enumerations several times, each time modulo a different prime. The enumerations were then reconstructed using the Chinese Remainder Theorem. Each run for a $`19\times 19`$ lattice took about 72 hours using 8 processors of a multiprocessor 1 GHz Compaq Alpha computer. Ten such runs were needed to uniquely specify the resultant numbers. Proceeding as above, we have calculated $`c_n(L)`$ for all $`n`$ for $`L17.`$ In other words, we have obtained the polynomials $`C_L(x)`$ for $`L17`$. In addition, we have computed $`C_{18}(1)`$ and $`C_{19}(1),`$ the total number of SAW crossing an $`18\times 18`$ and $`19\times 19`$ square respectively. We have also computed the corresponding quantities for cow-patch and transverse SAWs, denoted $`P_L(1)`$ and $`T_L(1)`$ respectively, for $`L19.`$ These are given in Table 2. In the question was asked whether $`C_{L,M}^{\frac{1}{LM}}`$ is decreasing in both $`L`$ and $`M`$? We can answer this in the negative, based on our enumerations. ## 4 Proofs of bounds Let $`𝒞(L)`$ be the set of self-avoiding walks crossing the $`L\times L`$ square from its south-west corner $`(0,0)`$ to its north-east corner $`(L,L)`$. Let $`C(L)`$ denote the cardinality of $`𝒞(L)`$. Let $`𝒯(L)`$ be the set of self-avoiding walks that traverse, the $`L\times L`$ square: by this, we mean that the walk starts from the west edge of the square and ends on the east edge (Figure 1). Let $`T(L)`$ be the cardinality of $`𝒯(L)`$. Finally, let $`𝒫(L)`$ be the set of cow-patches, of size $`L`$: a cow-patch is a configuration of mutually avoiding self-avoiding walks on the $`L\times L`$ square, such that each walk has both endpoints on the border of the square, but never contains an edge of the border (Figure 1). Let $`P(L)`$ be the number of cow-patches of size $`L`$. We first prove in this section that $$limC(L)^{1/L^2}=limT(L)^{1/L^2}=limP(L)^{1/L^2}=\lambda .$$ (7) Then, we prove the following bounds on $`\lambda `$: for $`L1`$, $$C(L)^{1/(L+1)^2}\lambda ,T(L)^{1/((L+1)(L+2))}\lambda ,\lambda (2P(L))^{1/L^2}.$$ Let us first focus on (7). As recalled in the previous sections, the convergence of $`C(L)^{1/L^2}`$ to $`\lambda `$ has been proved in earlier papers . For walks of $`𝒯(L)`$, a similar result follows from the fact that $$C(L)T(L)C(L+2).$$ The first inequality above is obvious. The second one is explained on the left of Figure 3. For cow-patches, the existence and value of $`limP(L)^{1/L^2}`$ follows from $$C(L1)P(L)C(L+3).$$ The first inequality is explained on the right of Figure 3. The second one is a bit more tricky. We borrow the following argument from . It is illustrated in Figure 4. Start from a cow-patch of size $`L`$. Colour all cells of the square in black and white, in such a way that the south-west corner of the square is black and each step included in one of the walks of the cow-patch is adjacent to a black cell and a white one. Surround the square by a layer of black cells, so as to obtain a square of size $`L+2`$, containing a certain number of white regions. For each white region, dig a tunnel (exactly one tunnel) in the outer layer to connect it to the outer world. In the figure thus obtained, the border of the black region forms a self-avoiding polygon, that includes each walk of the cow-patch. It remains to extend this polygon in a canonical way to obtain a walk of $`𝒞(L+3)`$, illustrated in the last panel of Figure 4. Let us now discuss lower and upper bounds on $`\lambda `$. The left-hand side of Figure 5 shows that for all $`\mathrm{}`$ and all odd $`k`$, it is possible to combine $`k^2`$ elements of $`𝒞(\mathrm{})`$ to form an element of $`𝒞(L)`$ with $`L=k(\mathrm{}+1)`$. In Figure 5, $`k=3`$. This shows that $$C(\mathrm{})^{k^2}C(L).$$ Hence $$C(\mathrm{})^{1/(\mathrm{}+1)^2}C(L)^{1/L^2}.$$ Taking the limit as $`k\mathrm{}`$ implies that for all $`\mathrm{}`$, $$C(\mathrm{})^{1/(\mathrm{}+1)^2}\lambda .$$ Similarly, let us try to pack transverse walks densely. The right-hand side of Figure 5 shows that for all $`\mathrm{}`$ and $`k`$, it is possible to combine $`k^2(\mathrm{}+1)(\mathrm{}+2)`$ elements of $`𝒯(\mathrm{})`$ to form an element of $`𝒞(L)`$ with $`L=k(\mathrm{}+1)(\mathrm{}+2)`$. This shows that $$T(\mathrm{})^{k^2(\mathrm{}+1)(\mathrm{}+2)}C(L).$$ Hence $$T(\mathrm{})^{1/((\mathrm{}+1)(\mathrm{}+2))}C(L)^{1/L^2}.$$ Taking the limit as $`k\mathrm{}`$ implies that for all $`\mathrm{}`$, $$\lambda T(\mathrm{})^{1/((\mathrm{}+1)(\mathrm{}+2))}.$$ (8) Let us finally give upper bounds for $`\lambda `$. Define a coloured cow-patch as a cow-patch in which the various regions are coloured in black and white, in such a way that two adjacent regions have different colours. Clearly, each cow-patch gives rise to 2 coloured cow-patches. Observe that there is a bijection between coloured cow-patches of size $`L`$ and the admissible, matrices of the same size, as defined in Section 5. Since an element of $`𝒞(L)`$, with $`L=k\mathrm{}`$, can be seen as the juxtaposition of $`k^2`$ admissible matrices (or coloured cow-patches) of size $`\mathrm{}`$, $$C(L)(2P(\mathrm{}))^{k^2}.$$ That is, $$C(L)^{1/L^2}(2P(\mathrm{}))^{1/\mathrm{}^2}$$ and by letting $`k\mathrm{}`$, we obtain Abbott and Hanson’s bound: for all $`\mathrm{}`$, $$\lambda (2P(\mathrm{}))^{1/\mathrm{}^2}.$$ One possible attempt to improve this bound is to consider generalised cow-patches, in which the walks are allowed to include edges lying on the west and south borders of the square (Figure 6). Let $`GP(L)`$ denote the number of generalised cow-patches of size $`L`$. Since an element of $`𝒞(L)`$, with $`L=k\mathrm{}`$, can be seen as the juxtaposition of $`k^2`$ generalised patches, the above argument gives $$\lambda GP(\mathrm{})^{1/\mathrm{}^2}.$$ We have not exploited this improvement, as it only changes the fourth significant digit of our bound. ## 5 Bounds on the growth constant $`\lambda `$ For the more general problem of SAW going from $`(0,0)`$ to $`(L,M)`$ on an $`L\times M`$ lattice, it was proved in that ###### Theorem 4 For each fixed $`M`$, $`lim_L\mathrm{}C_{L,M}^{\frac{1}{LM}}=\lambda _M`$ exists. Further, Abbott and Hanson state that a similar proof can be used to establish that $`lim_L\mathrm{}C_{L,L}^{\frac{1}{L^2}}:=\lambda `$ exists. This was proved rather differently in . ### 5.1 Upper bounds on $`\lambda `$ In an upper bound on the growth constant $`\lambda `$ was obtained by recasting the problem in a matrix setting. We give below an alternative method for establishing upper bounds, based on defining a superset of paths. We then show that these two methods are in fact identical. Following , consider any non-intersecting path crossing the $`L\times L`$ square. Label each unit square in the $`L\times L`$ lattice by 1 if it lies to the right of the path, and by 0 if it lies to the left. This provides a one-to-one correspondence between paths and a subset of $`L\times L`$ matrices with elements 0 or 1. Matrices corresponding to allowed paths are called admissible, otherwise they are inadmissible. Since the total number of $`L\times L`$ $`01`$ matrices is $`2^{L^2},`$ we immediately have the weak bound $`C_{L,L}2^{L^2}.`$ Of the 16 possible $`2\times 2`$ matrices, only 14 can correspond to portions of non-intersecting lattice paths. Note that there are only 12 actual paths from $`(0,0)`$ to $`(2,2)`$, but a further two matrices may correspond to paths that are embedded in a larger lattice. Thus we find the bound $`C_{L,L}14^{(L/2)^2},`$ so $`\lambda 1.9343..`$. Similarly, for $`3\times 3`$ lattices we find 320 admissible matrices (out of a possible 512), so $`\lambda 320^{1/9}=1.8982..`$ For $`4\times 4`$ lattices, claims that there are 22662 admissible matrices, but we believe the correct number to be 22816, giving the bound $`\lambda 1.8723..`$. We have made dramatic extensions of this work, using a combination of finite-lattice methods and transfer matrices, as described below, and have determined the number of admissible matrices up to $`19\times 19.`$ There are $`3.5465202\mathrm{}\times 10^{90}`$ such matrices, giving the bound $$\lambda 1.7817.$$ This bound is fully equivalent to the bound $`\lambda (2P_L)^{1/L^2}`$, where $`P_L`$ denotes the number of cow-patch configurations on the $`L\times L`$ lattice. This bound is proved below, in Section 4, and the equivalence follows upon colouring cow-patches by two colours, such that adjacent regions have different colours. Labeling the two colours $`0`$ and $`1`$ produces a $`01`$ matrix representation. ### 5.2 Lower bounds on $`\lambda `$ In the useful bound $$\lambda >\lambda _M^{\frac{M}{M+1}}$$ (9) is proved. The above evaluation of $`\lambda _2,`$ see (4), immediately yields $`\lambda >1.4892\mathrm{}`$. Based on exact enumeration, we have found the exact generating functions $`G_M(z)=_LC_{L,M}z^L`$ for $`M6.`$ For $`M=3`$ we find: $$G_3(z)=\frac{[1,4,4,36,39,26,50,6,15,1]}{[1,12,54,124,133,16,175,94,69,40,12,4,1]},$$ where we denote by $`[a_0,a_1,\mathrm{},a_n]`$ the polynomial $`a_0+a_1z+\mathrm{}+a_nz^n`$. As explained above, all the generating functions $`G_M(z)`$ are rational. For $`M=4,5,6`$, their numerator and denominators are found to have degree $`(26,27),(71,75)`$ and $`(186,186)`$ respectively, in an obvious notation. From these, we find the following values: $`\lambda _3=1.76331\mathrm{}`$, $`\lambda _4=1.75146\mathrm{}`$, $`\lambda _5=1.74875\mathrm{}`$ and $`\lambda _6=1.74728\mathrm{}`$. Then from eqn. (9) and $`\lambda _6`$ we obtain the bound $`\lambda >1.61339\mathrm{}`$. However, an alternative lower bound can be obtained from transverse SAWs, defined in Section 1. If $`T_L`$ denotes the number of transverse SAW on the $`L\times L`$ lattice, then we prove in the next section that $$\lambda T(L)^{1/((L+1)(L+2))}.$$ (10) From our enumerations of $`T(L)`$, given above for $`L19,`$ we obtain the improved bound $`\lambda >1.6284.`$ Combining our results for lower and upper bounds finally gives $$1.6284<\lambda <1.7817.$$ ## 6 Short walks crossing a square As defined in the introduction, let $`c_n(L)`$ be the number of $`n`$-step self-avoiding walks crossing an $`L\times L`$ square. Clearly, this number is zero when $`n`$ is odd and also when $`n<2L`$. It is almost as clear that $$c_{2L}(L)=\left(\genfrac{}{}{0pt}{}{2L}{L}\right).$$ Indeed, there are $`2L`$ steps in the path, of which $`L`$ must go north and $`L`$ must go east. Note that the number $`c_{2L}(L)`$ has asymptotic expansion $$\frac{4^L}{\sqrt{L\pi }}\left(1\frac{1}{4L}+\frac{1}{128L^2}+\frac{5}{1024L^3}+\mathrm{}\right).$$ Let us now prove that $$c_{2L+2}(L)=2L\left(\genfrac{}{}{0pt}{}{2L}{L2}\right).$$ A walk counted by $`c_{2L+2}(L)`$ has either $`L+2`$ vertical steps (and $`L`$ horizontal ones), or $`L`$ vertical steps (and $`L+2`$ horizontal ones). By symmetry, we can focus on the first case. Let $`w`$ be such a walk. We say that $`w`$ has a vertical defect., Among the $`L+2`$ vertical steps of $`w`$, exactly one goes south, while the $`L+1`$ others go north. The unique south step $`S`$ is necessarily preceded and followed by an east step, which we denote respectively $`E_1`$ and $`E_2`$. Let us mark $`E_1`$ and delete $`S`$ and $`E_2`$ (Figure 7). The marked path $`w^{}`$ thus obtained allows one to recover the original path $`w`$. It contains $`L+1`$ north steps and $`L1`$ east steps, one of which is marked. Moreover, the marked step cannot be at ordinate $`0`$, nor at ordinate $`L+1`$. Conversely, any walk $`w^{}`$ satisfying these properties is obtained (exactly once) from a walk counted by $`c_{2L+2}(L)`$ and having a vertical defect. The number of walks having $`L+1`$ north steps and $`L1`$ east steps is $`\left(\genfrac{}{}{0pt}{}{2L}{L1}\right)`$. Marking one of the east steps gives a factor $`(L1)`$. Now we must subtract the number of walks in which the marked step is either at level $`0`$ or at level $`N+1`$. Transforming the marked step into a vertical step shows that each of these two families of marked walks is in bijection with walks formed with $`L+2`$ up steps and $`L2`$ down steps. Putting these observations together gives $$c_{2L+2}(L)=2\left((L1)\left(\genfrac{}{}{0pt}{}{2L}{L1}\right)2\left(\genfrac{}{}{0pt}{}{2L}{L2}\right)\right)=2L\left(\genfrac{}{}{0pt}{}{2L}{L2}\right).$$ Note that the number $`c_{2L+2}(L)`$ has the asymptotic expansion $$\frac{L4^L}{\sqrt{L\pi }}\left(2\frac{33}{4L}+\frac{1345}{64L^2}\frac{23835}{512L^3}+\mathrm{}\right).$$ The same ideas may be used to find the value of $`c_{2L+4}(L)`$. We will prove that $$\frac{1}{2}c_{2L+4}(L)=\frac{(2L)!}{L!(L+4)!}\left(48+90L+8L^228L^33L^4+4L^5+L^6\right)2.$$ (11) First, note that $`c_{2L+4}(L)/2`$ is the number of self-avoiding walks (of length $`2L+4`$, crossing the $`L\times L`$ square) in which the first defect, that is, the first backward, step, is a south step. We focus on such walks, and study four distinct cases. The first three cases count walks having two south steps, and the last case counts walks having a south step and a west step (Figure 8). 1. The walk $`w`$ contains two adjacent south steps, $`S_1`$ and $`S_2`$. They are necessarily preceded by an east step $`E_1`$, and followed by another east step $`E_2`$. The walk has $`L+4`$ vertical steps and $`L`$ horizontal steps. Mark $`E_1`$, and delete $`S_1,S_2,E_2`$ in order to obtain a walk $`w^{}`$ with $`L+2`$ north steps and $`L1`$ east steps, one of which is marked. In $`w^{}`$, the marked step cannot be at level $`0,1,L+1`$ or $`L+2`$. Using the same ingredients as above, we obtain the number of such walks as $$(L1)\left(\genfrac{}{}{0pt}{}{2L+1}{L1}\right)4\left(\genfrac{}{}{0pt}{}{2L+1}{L2}\right).$$ 2. The walk contains a sequence $`E_1S_1E_2S_2E_3`$. Again, $`w`$ has $`L+4`$ vertical steps and $`L`$ horizontal steps. Mark $`E_1`$, and delete $`S_1,E_2,S_2`$ and $`E_3`$ in order to obtain a walk with $`L+2`$ north steps and $`L2`$ east steps, one of which is marked. In $`w^{}`$, the marked step cannot be at level $`0,1,L+1`$ or $`L+2`$. The number of such walks is $$(L2)\left(\genfrac{}{}{0pt}{}{2L}{L2}\right)4\left(\genfrac{}{}{0pt}{}{2L}{L3}\right).$$ 3. The walk contains a sequence $`E_1S_1E_2`$, and, further away, another sequence $`E_3S_2E_4`$, disjoint from the first one. Again, $`w`$ has $`L+4`$ vertical steps and $`L`$ horizontal steps. Mark the steps $`E_1`$ and $`E_3`$, delete $`S_1,E_2,S_2`$ and $`E_4`$ in order to obtain a walk with $`L+2`$ north steps and $`L2`$ east steps, two of which are marked. Note that, in $`w^{}`$, the first marked step cannot lie at level $`0,L+1`$ or $`L+2`$, while the second marked step cannot lie at level $`0,1`$ or $`L+2`$. Using the same ingredients as above, combined with the inclusion-exclusion principle, we find the number of such walks as $$\left(\genfrac{}{}{0pt}{}{L2}{2}\right)\left(\genfrac{}{}{0pt}{}{2L}{L2}\right)2\left[(L3)\left(\genfrac{}{}{0pt}{}{2L}{L3}\right)\left(\genfrac{}{}{0pt}{}{2L}{L4}\right)\right]4\left(\genfrac{}{}{0pt}{}{2L}{L4}\right)2\left(\genfrac{}{}{0pt}{}{2L}{L4}\right)+5\left(\genfrac{}{}{0pt}{}{2L}{L4}\right)$$ $$=\left(\genfrac{}{}{0pt}{}{L2}{2}\right)\left(\genfrac{}{}{0pt}{}{2L}{L2}\right)2(L3)\left(\genfrac{}{}{0pt}{}{2L}{L3}\right)+\left(\genfrac{}{}{0pt}{}{2L}{L4}\right).$$ 4. The walk $`w`$ contains a sequence $`E_1S_1E_2`$, and, further away, a sequence $`N_1W_1N_2`$ (with obvious notations). It thus contains $`L+2`$ vertical steps and $`L+2`$ horizontal ones. Mark the steps $`E_1`$ and $`N_1`$, delete $`S_1,E_2,W_1`$ and $`N_2`$ in order to obtain a walk $`w^{}`$ with $`L`$ north steps and $`L`$ east steps, in which one step of each type is marked in such a way that the east marked step comes before the north marked step. In $`w^{}`$, the two marked steps cannot be consecutive (or $`w`$ would not be self-avoiding), the east marked step cannot lie at level $`0`$, and the north marked step cannot lie at abscissa $`L`$. Again, the inclusion-exclusion principle applies and gives the number of such walks as $$\frac{1}{2}L^2\left(\genfrac{}{}{0pt}{}{2L}{L}\right)(2L1)\left(\genfrac{}{}{0pt}{}{2L2}{L1}\right)2L\left(\genfrac{}{}{0pt}{}{2L}{L1}\right)+2\left(\genfrac{}{}{0pt}{}{2L1}{L1}\right)+\left[\left(\genfrac{}{}{0pt}{}{2L}{L}\right)1\right]1.$$ Putting together the four partial results we have obtained gives (11). Note that the number $`c_{2L+4}(L)`$ has the asymptotic expansion $$\frac{L^24^L}{\sqrt{L\pi }}\left(2\frac{49}{4L}+\frac{2913}{64L^2}\frac{92971}{512L^3}+\mathrm{}\right).$$ The above argument suggests that it is very likely that, for every fixed $`K`$, the sequence $`c_{2L+2K}(L)`$, for $`L0`$, is polynomially recursive \[16, 17, Ch. 6\]. While it would probably be possible to find the number of possible paths of length $`2L+6`$, the number of special cases that must be treated would become onerous. We have therefore resorted to a numerical study for walks of length $`2L+2K,`$ $`K>2`$, based on our enumerations. For $`K=3`$ we found $$\frac{L^34^L}{\sqrt{L\pi }}\left(\frac{4}{3}\frac{49}{6L}+\frac{1931\pm 1}{64L^2}+\mathrm{}\right),$$ while the corresponding result for $`K=4`$ is $$\frac{L^44^L}{\sqrt{L\pi }}\left(\frac{2}{3}+\frac{11}{4L}+\mathrm{}\right).$$ We can give a heuristic argument for the general form of the leading term in the asymptotic expansion of the number of walks of length $`2L+2K`$ which gives as the leading order term $`\frac{4^L}{\sqrt{L\pi }}\frac{(2L)^K}{K!}.`$ Here the first factor is given by the number of ways of choosing the backbone, $`\left(\genfrac{}{}{0pt}{}{2L}{L}\right)\frac{4^L}{\sqrt{L\pi }}`$ and the second is given by the number of ways of placing $`K`$ defects (or backward steps) on a path of length $`2L`$, which is just $`(2L)^K`$. The defects are indistinguishable, introducing the factor $`K!`$. This argument can be refined into a proof, for $`K=\mathrm{o}(L^{1/3})`$ by following the steps, mutatis mutandis in the proof of a similar result given in . ## 7 Numerical analysis It has been proved that $`lim_L\mathrm{}C_{L,L}^{\frac{1}{L^2}}=\lambda `$ exists. From this it is likely that $`R_L=C_{L+1,L+1}/C_{L,L}\lambda ^{2L}`$ though this has not been proved. Accepting this, the generating function $`(x)=_LR_Lx^L`$ will have radius of convergence $`x_c=1/\lambda ^2`$, which we can estimate accurately using differential approximants . In this way we estimate that for the crossing problem $`x_c=0.32858(5)`$, for the transverse problem $`x_c=0.3282(6)`$ and for the cow-patch problem $`x_c=0.328574(2)`$. It is reassuring to see, from our numerical studies, that $`\lambda `$ appears to be the same for the three problems, as proved above, and we estimate that $`\lambda =1.744550(5)`$. We now speculate on the sub-dominant terms. For SAW on an infinite lattice, it is widely accepted that $`c_nconst.\mu ^nn^g`$ where $`c_n`$ is the number of $`n`$ step SAW equivalent up to a translation. It seems at least a plausible speculation that, for SAW crossing an $`L\times L`$ lattice, the number going from $`(0,0)`$ to $`(L,L)`$ is given by $`A\lambda ^{L^2+bL}L^\alpha .`$ We have investigated this possibility numerically, and found it to be supported by the data, to some extent. We fitted the data to the assumed form, fixing the value of $`\lambda `$ at our best estimate, 1.744550. This then leaves two unknown parameters $`b`$ and $`\alpha .`$ For cow-patch walks we find $`b0.8558`$ and $`\alpha 0.500.`$ This suggests asymptotic behaviour $`A_P\lambda ^{L^2+0.8558L}/\sqrt{L}`$, and we estimate $`A_P0.52`$. For transverse walks and walks crossing a square $`b`$ is quite small, most likely zero. A value of $`b=0`$ would imply the absence of a term O$`(\lambda ^{bL})`$, or possibly the presence of a term O$`(\mathrm{log}L)`$, or some power of a logarithm. We have investigated the latter possibility by including a logarithmic factor, and found that the data does not support the presence of such a term for either class of walk. Of course, we cannot rule out some small power of a logarithm, but this seems less likely than the absence of a term O$`(\lambda ^{bL})`$. We next investigated the possibility that the subdominant term is O$`(L^\alpha ).`$ A simple ratio analysis then led to the estimates $`\alpha =0.7`$ for walks crossing a square, and $`\alpha =1.0`$ for transverse walks. If our assumed form is correct, we expect these estimates to be accurate to within 10-15%. We also studied the sequence whose terms are given by the quotient $`T_L/C_L.`$ This has the advantage that the $`\lambda `$ dependence cancels, and so our result is independent of any uncertainty in the value of $`\lambda .`$ We find that $`T_L/C_Lconst.L^{1.7}`$ This is in agreement with the estimates of $`\alpha `$ found separately, for the two series. Thus we very tentatively speculate that $`C_L8\lambda ^{L^2}/L^{0.7}`$ and $`T_L9\lambda ^{L^2}L,`$ where the amplitude estimates follow by the simple expedient of fitting the assumed $`L`$ dependent form to the data, term-by-term, and extrapolating the resulting sequence of amplitude estimates. Given the sensitivity of the amplitudes to both $`\lambda `$ and $`\alpha `$, we do not feel confident quoting an uncertainty for the amplitudes. Whittington and Guttmann and later Burkhardt and Guim studied the behaviour of the mean number of steps in a path on an $`L\times L`$ lattice $$n(x,L)=\frac{_nnc_n(L)x^n}{_nc_n(L)x^n}$$ (12) as well as the fluctuations of this quantity $$V(x,L)=\frac{_nn^2c_n(L)x^n}{_nc_n(L)x^n}n(x,L)^2$$ (13) which is a kind of heat capacity. As discussed above, a phase transition takes place as one varies the fugacity $`x`$ associated with the walk length. At a critical value $`x_c`$, the average walk length of a path on an $`L\times L`$ lattice changes from $`\mathrm{\Theta }(L)`$ to $`\mathrm{\Theta }(L^2).`$ In the critical fugacity was proved to satisfy $`1/\mu x_c\mu _H`$, where $`\mu _H`$ is the growth constant for Hamiltonian SAW on the square lattice, and on the basis of numerical studies conjectured to be $`x_c=1/\mu `$ exactly. In the conjecture was proved. Here we also study the behaviour at $`x=x_c`$ and find that $`n(x,L)=\mathrm{\Theta }(L^{1/\nu })`$ where the numerical evidence is consistent with $`\nu =3/4`$. Similar conclusions were reached earlier in . For any given value of $`L`$ the fluctuation $`V(x,L)`$ is observed to have a single maximum located at $`x_c(L)`$ (see top left panel of Figure 9). We study in detail the behaviour of $`V(x,L)`$, which we expect to obey a standard finite-size scaling Ansatz $$V(x,L)L^{2/\nu }\stackrel{~}{V}((xx_c)L^{1/\nu }),$$ (14) (which is equivalent to (2) of ) where $`\stackrel{~}{V}(y)`$ is a scaling function. From this it follows that the position and the height of the peak in $`V(x,L)`$ scale as $`x_c(L)x_cL^{1/\nu }`$ and $`V_{\mathrm{max}}(L)L^{2/\nu }`$. In table 3 we have listed the numerical values of the mean-length at $`x_c`$ and the position and height of the maximum of the fluctuations. We analyse this data by forming the associated generating functions, $`N(z)=_Ln(x,L)z^L`$ etc., and using differential approximants. Given the expected asymptotic behaviour of these quantities the generating functions should have a singularity at $`z_c=1`$ with critical exponents $`1/\nu 1`$ (average length at $`x_c`$), $`1/\nu 1`$ (position of the peak), and $`2/\nu 1`$ (height of the peak). In table 4 we list the results from an analysis of the generating functions using second order differential approximants. The estimates for the exponents are not very accurate (which is not surprising given the short length of the series) but are fully consistent with $`\nu =3/4`$. Finally, in Figure 9 we perform a more detailed analysis to confirm the conjectured scaling form for $`V(x,L)`$. In the top left panel we have simply plotted $`V(x,L)`$ as a function of the fugacity $`x`$ to confirm the single peak behaviour. In the top right panel we have plotted $`x_c(L)`$ and $`V_{\mathrm{max}}`$ vs. $`L`$ in a log-log plot, thus confirming that these quantities grows as a power-law with $`L`$ (the straight lines, drawn as a guide to the eye, have slopes $`1/\nu =4/3`$ and $`2/\nu =8/3`$, respectively). In the bottom panels we check numerically the scaling Ansatz for $`V(x,L)`$. In the left panel we plot $`V(x,L)/L^{8/3}`$ vs. the scaling variable $`(xx_c)L^{4/3}`$ obtaining a reasonable scaling collapse. A better idea of the quality of the scaling collapse can be gauged from the plot in the bottom right panel. Here we plot the difference between consecutive scaling plots from the left panel. More precisely we plot $`D(x,L)=V(x,L)/L^{8/3}V(x^{},L1)/(L1)^{8/3}`$ vs. $`(xx_c)L^{4/3}`$, where $`x^{}`$ is chosen so that the scaled variables coincide, e.g., $`(xx_c)L^{4/3}=(x^{}x_c)(L1)^{4/3}`$. ## 8 Walks crossing the square and hitting the centre In Knuth also considered the problem of self-avoiding walks crossing the square and passing through the centre $`(L/2,L/2)`$ of the grid (with $`L`$ being even). Denote the number of such walks by $`c(L).`$ Then a straightforward variant of the method of proof used in Section 4 can be applied to prove that $$\underset{L\mathrm{}}{lim}c(L)^{1/L^2}=\lambda ^2.$$ Knuth used Monte Carlo simulations to estimate the fraction of paths hitting the centre point and found for $`L=10`$ that $`81\pm 10`$ percent of all paths do hit the centre. He then went on to say that “perhaps nobody will ever know the true answer.” Naturally we cannot let Knuth’s challenge go unanswered. It is very simple to modify the transfer-matrix algorithm to ensure that all paths pass through a given vertex. We just make sure that when we do the updating at the given vertex the input state $`{}_{}{}^{}00_{}^{}`$ (no occupied incoming edges) has only one output state $`{}_{}{}^{}12_{}^{}`$, while the output $`{}_{}{}^{}00_{}^{}`$ (no outgoing occupied edges) is disallowed at this vertex. We can thus answer Knuth’s query and state for all to know that for $`L=10`$ a fraction $`1243982213040307428318660/1568758030464750013214100=0.792972\mathrm{}`$ of all paths pass through the centre. In Table 5 we have listed the number of paths passing through the centre for $`L18`$. The fact that $`C(L)/c(L)`$ appears to be going to a constant implies that not only is the asymptotically dominant behaviour of both $`C(L)`$ and $`c(L)`$ the same, but so must the sub-dominant behaviour. We note the useful mnemonic that the ratio appears close to $`\sqrt{\pi /5}=0.79266\mathrm{},`$ though we have no idea how to prove or disprove that this is the correct value. ## 9 Hamiltonian walks Hamiltonian walks can only exist on $`2L\times 2L`$ lattices. For lattices with an odd number of edges, one site must be missed. A Hamiltonian walk is of length $`4L(L+1)`$ on a $`2L\times 2L`$ lattice. The number of such walks grows as $`\tau ^{4L^2},`$ where we find $`\tau 1.472`$ based on exact enumeration up to $`17\times 17`$ lattices. In Jacobsen and Kondev gave a field-theoretical estimate of the growth constant for Hamiltonian SAW on the square lattice as $`1.472801\pm 0.00001`$. These were walks confined to a square geometry, but not restricted as to starting and end-points as are those we consider here. Nevertheless, it seems likely that we are estimating the same quantity, so our results can be seen as providing support for the view that the field theory is estimating precisely the same quantity as our enumerations. That is to say, this appears to be precisely the same as the corresponding result for Hamiltonian walks on an $`L\times L`$ lattice, in the large $`L`$ limit. These estimates are about $`20\%`$ less than $`\lambda ,`$ the growth constant for all paths. In it is proved that $`2^{1/3}\tau 12^{1/4}.`$ Numerically this evaluates to $`1.260\tau 1.861.`$ We can improve on these bounds as follows: we define cow-patch walks to be Hamiltonian if every vertex of the square not belonging to the border of the square belongs to one of the SAWs of the cow-patch. Then the upper bounds given above translate verbatim into upper bounds for $`\tau ,`$ while lower bounds are given by Hamiltonian traversing paths and eqn. (8). In this way we find $`1.429<\tau <1.530.`$ As we have shown above that $`1.6284<\lambda ,`$ this proves that $`\tau <\lambda .`$ The number of Hamiltonian paths $`H_L`$ for $`L`$ even, and paths that visit all but one site, for $`L`$ odd, are given in Table 6. The number of Hamiltonian cow-patch paths $`HP_L`$ for $`L`$ even, and cow-patch paths that visit all but one site, for $`L`$ odd, are given in Table 7. The number of Hamiltonian transverse paths $`HT_L`$ for $`L`$ even, and transverse paths that visit all but one site, for $`L`$ odd, are given in Table 8. ## E-mail or WWW retrieval of series The series for the problems studied in this paper are available by request from I.Jensen@ms.unimelb.edu.au or via the world wide web http://www.ms.unimelb.edu.au/~iwan/ by following the relevant links. ## Acknowledgments We would like to thank Stu Whittington for helpful comments on the manuscript. MBM was partially supported by the European Commission’s IHRP Programme, grant HPRN-CT-2001-00272, “Algebraic Combinatorics in Europe”. Two of us (AJG and IJ) have been supported by grants from the Australian Research Council. The calculations presented in this paper were performed on facilities provided by the Australian Partnership for Advanced Computing (APAC) and the Victorian Partnership for Advanced Computing (VPAC).
warning/0506/cond-mat0506205.html
ar5iv
text
# Microscopic theory of surface-enhanced Raman scattering in noble-metal nanoparticles ## I Introduction Surface-enhanced Raman scattering (SERS) has been one of the highlights of optical spectroscopy in metal nanostructures during past 25 years. schatz-review02 Recent interest in SERS stems from the discovery of extremely strong single-molecule SERS in silver nanoparticle aggregates, nie-sci97 ; kneipp-prl97 as well as from nanoparticle-based applications such as, e.g., biosensors mirkin-science02 that rely on sensitivity of SERS to small concentrations of target molecules. The main mechanism of SERS has long been known as electromagnetic (EM) enhancement schatz-review02 ; moskovits-rmp85 ; kerker-ao80 ; gersten-jcp81 of dipole moment of a molecule by the strong local field of surface plasmon (SP) resonance in a nanoparticle. EM mechanism is especially effective when a cluster of nanoparticles is concentrated in a small region (“hot spot”). kneipp-cr99 ; brus-jpcb00 ; moskovits-tap02 ; rothberg-pnas04 A combined effect of SP local fields from different particles acting on a molecule trapped in a gap can result in a giant (up to $`10^{14}`$) enhancement of the Raman scattering crossection. stockman-prb96 ; markel-prb96 ; kall-prl99 ; kall-prb00 ; corni-jcp02 ; stockman-prl03 Other mechanisms contributing to SERS can involve electron tunneling between a molecule and a nanoparticle. otto-jpcm92 The conventional description of EM enhancement is based on classical Mie scattering theory. kerker-ao80 ; gersten-jcp81 The dipole moment of a molecule at distance $`𝐫_0`$ form a particle center is enhanced by a factor $`\alpha _p(\omega )/r_0^3`$, where $`\alpha _p=R^3\left[(ϵ1)/(ϵ+2)\right]`$ is the particle polarizability, $`R`$ is its radius, and $`ϵ(\omega )`$ is metal dielectric function. The far-field of molecular dipole, radiating at Stokes-shifted frequency $`\omega _s`$, is, in turn, comprised of direct and Mie-scattered fields. The latter contributes another factor $`\alpha _p(\omega _s)/r_0^3`$, so that the total field enhancement is $`\alpha _p(\omega )\alpha _p(\omega _s)/r_0^6`$ and Raman crossection is proportional to $`|\alpha _p|^4/r_0^{12}`$. At frequencies close to the SP pole in $`\alpha _p`$, this enhancement can reach $`10^6`$. Note that, within classical description, the dependence of SERS on nanoparticle size, coming from geometrical factor in $`\alpha `$, is weak if the molecule is sufficiently close to nanoparticle surface. The classical approach is valid for relatively large nanoparticles, where the effect of confining potential on electronic states is negligible. For nanoparticle sizes $`R5`$ nm, the lifetime of SP is reduced due to the Landau damping by single-particle excitations accompanied by momentum transfer to the surface.kubo-jpsj66 This results in a broadening of SP resonance peak by the amount of level spacing at the Fermi energy, $`\gamma _sv_F/R`$ ($`v_F`$ is the Fermi velocity), and in the corresponding decrease of the SP field amplitude. For not very small nanoparticles, the effect of Landau damping on SERS can be treated semiclassically schatz-review02 by incorporating the quantum-size correction $`\gamma _s`$ in the Drude dielectric function of metal. Recent resonance fluorescence measurements on small gold particles, feldmann-prl02 ; feldmann-nl05 ; nie-jacs02 however, indicate strong deviations of the plasmon-induced enhancement from that predicted by semiclassical models. gersten-jcp81 For clusters with electron number $`N<100`$, SP lifetime is reduced to several fs and SERS is diminished. Note that some enhancement of Raman signal due to single-particle resonances remains even in small clusters containing several atoms. dickson-prl05 In this paper, we study SERS for small nanoparticles several nanometers in diameter, i.e., in the intermediate regime between classical particles and small clusters. Our chief observation is that, in this crossover regime, SERS magnitude is determined by interplay between several competing quantum-size effects, including the aforementioned SP Landau damping as well as the modification of electron screening in the surface region. The latter produces an opposite trend towards a relative increase of SERS when the molecule is located in a close proximity to the metal surface. The underlying mechanism is related to different effects that the confining potential has on d-band and sp-band electron states. Namely, the deviation of potential well from the rectangular shape gives rise to a larger effective radius for the higher-energy sp-electrons. persson-prb85 This effect is further amplified by the spillover of sp-band electron density, due to tunneling into the barrier, as contrasted to the essentially step-like density profile of localized d-electrons. As a result, in a surface layer of thickness $`12`$ Å, the d-electron population is diminished and, hence, the interband (i.e. due to d–sp transitions) polarizability is strongly reduced. liebsch-prb93 ; liebsch-prb95 The resulting reduction of screening in the surface region leads to a greater strength of electron-electron interactions in the sp-band that was observed, e.g., in a faster, as compared to bulk metal, electron relaxation in Ag nanoparticles measured using ultrafast pump-probe spectroscopy. voisin-prl00 It is, therefore, natural to expect that such underscreening should lead to additional enhancement of local field outside of the nanoparticle. Note, however, that here the effect of spillover is twofold: while it leads to an increase of volume fraction of underscreened region and hence to stronger local field, especially in small nanoparticles, it can also, by itself, have an opposite effect by smearing out the otherwise sharp classical boundary. dignam-jcp92 Thus, in small nanoparticles, the enhancement magnitude is determined by a delicate interplay of competing quantum-size effects, and must, therefore, be described within a consistent microscopic approach. Such an approach, based on time-dependent local density approximation (TDLDA), ekardt-prb85 is developed in this paper. The paper is organized as follows. In Section II, we derive the general expression for polarizability of molecule-nanoparticle system. In Section III, we derive a self-consistent system for local potential that determines the enhancement factor. In Section IV we present the results of our numerical calculations. Section V concludes the paper. ## II Polarizability of molecule-nanoparticle system We start with formulating SERS in terms of quantum transitions in the interacting molecule-nanoparticle system. We assume that the molecule is located at distance $`r_0`$ from the nanoparticle center and that the overall system size is much smaller than radiation wavelength so the retardation effects can be ignored. kreibig-book In the absence of direct electron tunneling, otto-jpcm92 the interactions within excited molecule-nanoparticle system are caused by nonradiative transitions accompanied by energy transfer between a molecule and a nanoparticle, similar to Forster transfer in two-molecule systems. lakowicz-book Namely, an electron-hole pair can nonradiatively recombine by transferring its energy to SP (and vice versa) via dynamically-screened Coulomb interaction. shahbazyan-prl98 Feynman diagrams of processes contributing to polarizability of molecule-nanoparticle system, $`\stackrel{~}{\alpha }`$, are shown in Fig. 1: (a) incident photon with energy $`\omega `$ is absorbed by the molecule and reemitted with Stokes-shifted energy $`\omega _s`$; (b) after absorbing a photon, excited molecule nonradiatively recombines transferring its energy to SP in the nanoparticle, which emits a photon; (c) SP, excited by incident light, transfers its energy to the molecule, which emits a photon; and (d) after energy transfer from SP to molecule, the latter transfers the energy back to SP, which emits a photon. Correspondingly, the system polarizability is (in operator form) $`\stackrel{~}{\alpha }=\alpha +\alpha U\mathrm{\Pi }+\mathrm{\Pi }U\alpha +\mathrm{\Pi }U\alpha U\mathrm{\Pi },`$ (1) where $`\alpha `$ is molecular polarizability, $`\mathrm{\Pi }`$ is density-density response function of a nanoparticle in medium, and $`U`$ is the Coulomb potential. The Raman polarizability is obtained by calculating the matrix element of $`\stackrel{~}{\alpha }`$ between incoming and outgoing photon states with energies $`\omega `$ and $`\omega _s`$, respectively. The interaction of the molecule with the nanoparticle involves matrix elements $`e|\varphi (𝐫)|g`$, where $`|g`$ and $`|e`$ stand for the molecule ground and excited electronic bands, respectively, and $`\varphi (\omega ,𝐫)={\displaystyle 𝑑𝐫_1𝑑𝐫_2U(𝐫𝐫_1)\mathrm{\Pi }(\omega ,𝐫_1,𝐫_2)\varphi _0(𝐫_2)}`$ (2) is the nanoparticle response to external photon potential, $`\varphi _0(𝐫)`$. Since the length scale of $`\varphi (𝐫)`$ is much larger than the molecule size, we have $`e|\varphi (𝐫)|g𝝁\varphi (𝐫_0)`$, where $`𝝁`$ is the dipole matrix element of corresponding molecular transition. long-book The averaging over random orientations of $`𝝁`$ can be accounted by assuming isotropic Raman polarizability tensor $`\alpha `$. The nanoparticle contribution then factors out, $`\stackrel{~}{\alpha }=\alpha M`$, with $`M=1+{\displaystyle \frac{1}{E_0^2}}[𝐄_0\varphi (\omega ,𝐫_0)+𝐄_0\varphi (\omega _s,𝐫_0)`$ $`+\varphi (\omega ,𝐫_0)\varphi (\omega _s,𝐫_0)],`$ (3) where $`𝐄_0`$ is the electric field in the absence of nanoparticle. For incident field, $`𝐄_i`$, polarized along the $`z`$-axis, $`\varphi _0=eE_ir\mathrm{cos}\theta `$, we have $`𝐄_0=𝐄_i/ϵ_m`$, where $`ϵ_m`$ is the dielectric constant of medium. ## III Enhancement factor To evaluate the local potential $`\varphi (\omega ,𝐫)`$ within TDLDA approach, we present it in the form $`\varphi (\omega ,𝐫)=e^2{\displaystyle d^3r^{}\frac{\delta n(\omega ,𝐫^{})}{|𝐫𝐫^{}|}},`$ (4) where the induced density, $$\delta n(𝐫)=𝑑𝐫^{}\mathrm{\Pi }(𝐫,𝐫^{})\varphi _0(𝐫^{})=\delta n_s(𝐫)+\delta n_d(𝐫)+\delta n_m(𝐫),$$ (5) contains contributions from sp-electrons, d-electrons, and surrounding medium, respectively (hereafter we suppress frequency dependence). We adopt the two-region model that combines a quantum-mechanical description for sp-band electrons and phenomenological treatment d-electrons with bulk-like ground-state density $`n_d`$ in the region confined by $`R_d<R`$. liebsch-prb93 This model has been used for calculations of polarizabilities of small Ag nanoparticles and clusters, kresin-prb95 ; lerme-prl98 but it remains reliable for relatively large electron numbers, $`N>1000`$. The induced density of sp-band electrons is determined from TDLDA equation $`\delta n_s(𝐫)={\displaystyle d^3r^{}P_s(𝐫,𝐫^{})\left[\mathrm{\Phi }(𝐫^{})+V_x^{}[n(r^{})]\delta n_s(𝐫^{})\right]},`$ (6) where $`\mathrm{\Phi }=\varphi _0+\varphi `$ is the full potential, $`P_s(𝐫,𝐫^{})`$ is the polarization operator for noninteracting sp-electrons, $`V_x^{}[n(r^{})]`$ is the (functional) derivative of the exchange-correlation potential and $`n(r)`$ is the ground-state electron density. The latter is obtained in a standard way by solving Kohn-Sham equations. To close the system, we need to express the full potential $`\mathrm{\Phi }(𝐫)`$ via $`\delta n_s(𝐫)`$. This is accomplished be relating $`\delta n_d(𝐫)`$ and $`\delta n_m(𝐫)`$ back to $`\mathrm{\Phi }(𝐫)`$ as $`e^2\delta n_d(𝐫)=\left[\chi _d(r)\mathrm{\Phi }(𝐫)\right],`$ $`e^2\delta n_m(𝐫)=\left[\chi _m(r)\mathrm{\Phi }(𝐫)\right],`$ (7) where $`\chi _d(r)=\left[(ϵ_d1)/4\pi \right]\theta (R_dr)`$ is the interband susceptibility with the step function enforcing the boundary conditions and, correspondingly, $`\chi _m(r)=\left[(ϵ_m1)/4\pi \right]\theta (rR)`$ is the susceptibility of surrounding medium. The derivation is given in Appendix. The final result is conveniently expressed in terms of expansion of $`\mathrm{\Phi }(𝐫)`$ and $`\delta n(𝐫)`$ in spherical harmonics. For non-resonant Raman scattering, only dipole ($`L=1`$) terms contribute to the local field. The final expressions for dipole component $`\mathrm{\Phi }(r)`$ can be presented as a decomposition (see the Appendix) $$\mathrm{\Phi }=\frac{1}{ϵ(r)}\left[\varphi _0(r)+\delta \varphi _d(r)+\delta \varphi _s(r)\right],$$ (8) where $`\varphi _0(r)=eE_ir`$, $`\delta \varphi _d(r)=\beta (r/R)\varphi _0(R)\lambda _m(1a^3\lambda _d)/\eta `$ $`\beta (r/R_d)\varphi _0(R_d)\lambda _d(12\lambda _m)/\eta `$ (9) and $`\delta \varphi _s(r)={\displaystyle 𝑑r^{}r^2K(r,r^{})\delta n_s(r^{})}.`$ (10) Here $`ϵ(r)=(ϵ_d`$, 1, $`ϵ_m`$) for $`r`$ in the intervals \[$`(0,R)`$, $`(R_d,R)`$, $`(R,\mathrm{})`$\], respectively, and $`\lambda _d={\displaystyle \frac{ϵ_d1}{ϵ_d+2}},\lambda _m={\displaystyle \frac{ϵ_m1}{2ϵ_m+1}},\eta =12a^3\lambda _d\lambda _m,`$ (11) with $`a=R_d/R`$. The kernel $`K(r,r^{})`$, relating the induced potential and density of sp-electrons, is given by $`K(r,r^{})=u(r,r^{})`$ $`\beta (r/R_d)\left[u(R_d,r^{})2a\lambda _mu(R,r^{})\right]\lambda _d/\eta `$ $`+\beta (r/R)\left[u(R,r^{})a^2\lambda _du(R_d,r^{})\right]\lambda _m/\eta ,`$ (12) where $`u(r,r^{})=4\pi r_</3r_>^2`$ is the dipole term of Coulomb potential expansion and $`\beta (x)=x^2\theta (x1)2x\theta (1x)`$. With decomposition (8), the TDLDA equation (6) takes the form $`\delta n_s(r)={\displaystyle 𝑑r^{}r^2P_s(r,r^{})\frac{1}{ϵ(r^{})}\left[\varphi _0(r^{})+\delta \varphi _d(r^{})\right]}`$ $`+{\displaystyle }dr^{}r^2P_s(r,r^{}){\displaystyle \frac{1}{ϵ(r^{})}}[{\displaystyle }dr^{\prime \prime }r^{\prime \prime 2}K(r^{},r^{\prime \prime })\delta n_s(r^{\prime \prime })`$ $`+V_x^{}(r^{})\delta n_s(r^{})],`$ (13) Note that $`\mathrm{\Phi }(r)`$ is continuous at $`r=R_d,R`$. Equations (8)–(III) determine self-consistently the spatial distribution of local potential near small noble-metal nanoparticles. Here $`\delta \varphi _d(r)`$ is the induced potential due to d-electrons and surrounding medium. Their effect on the sp-electron potential, $`\delta \varphi _s(r)`$, is encoded in the kernel $`K(r,r^{})`$. For $`ϵ_d=ϵ_m=1`$, we have $`K(r,r^{})=u(r,r^{})`$ and $`\delta \varphi _d(r)=0`$, recovering the case of simple metal particles in vacuum. If the molecule is not too close to the surface ($`d1`$Å), i.e., there is no significant overlap between molecular orbitals and electronic states, then Eqs. (4) and (8)–(III) yield $$\delta \mathrm{\Phi }(r_0)\frac{1}{ϵ(r_0)}\left[\delta \varphi _d(r_0)+\delta \varphi _s(r_0)\right]=\frac{eE_i}{ϵ_mr_0^2}\alpha _p,$$ (14) where $`\alpha _p(\omega )`$ is the nanoparticle polarizability. The expression for $`\alpha _p`$ is given in Appendix. Field enhancement coefficient $`M`$, Eq. (II), then takes the form $`M=1+(1+\mathrm{cos}^2\theta _0){\displaystyle \frac{\alpha _p(\omega )+\alpha _p(\omega _s)}{r_0^3}}`$ $`+(1+3\mathrm{cos}^2\theta _0){\displaystyle \frac{\alpha _p(\omega )\alpha _p(\omega _s)}{r_0^6}}.`$ (15) In this case, enhancement retains the same functional dependence on particle polarizability as in classical theory;kerker-ao80 however, $`\alpha _p`$ is now determined microscopically. ## IV Numerical results and discussion Below we present our results for SERS enhancement factor $`|M|^2`$ for Ag nanoparticles in a medium with dielectric constant $`ϵ_m=1.5`$. Calculations were carried for number of electrons ranging from $`N=92`$ to $`N=3028`$, corresponding to particle diameters in the range $`D1.44.5`$ nm. To ensure spherical symmetry, only closed-shell“magic numbers” were used. koch-prb96 For such sizes, the Ag band-structure remains intact. The ground state energy spectrum and wave-functions were obtained by solving the Kohn-Sham equations for jellium model ekardt-prb85 with the Gunnarsson-Lundqvist exchange-correlation potential; lundqvist-prb77 the interaction strength was appropriately modified to account for static d-band screening. The ground state density of sp-band electrons, $`n(r)`$, exhibits characteristic Friedel oscillations, while the spatial extent of spillover is $`2`$ a.u. (see Fig. 2). These results were used as input in the numerical solution of TDLDA system (8)–(III). The molecule was located along the $`z`$-axis ($`\theta _0=0`$) at a distance $`d=5`$ a.u. from effective boundary with radius $`R=r_sN^{1/3}`$, where $`r_s=(4\pi n/3)^{1/3}`$ ($`r_s=3.0`$ a.u. for Ag), ensuring no direct overlap with the nanoparticle. In the calculation of optical response, the experimental data for $`ϵ_d(\omega )`$ in Ag was used palik-book . We also assumed that the Stokes shift is much smaller than $`\gamma _sv_F/R`$ ($`\gamma _s0.5`$ eV for $`D=3.0`$ nm) and ignored the difference between $`\omega `$ and $`\omega _s`$. The calculated absorption spectra for different nanoparticle sizes are shown in Fig. 3. For medium dielectric constant $`ϵ_m=1.5`$, the position of SP resonance at $`3.2`$ eV is well below of the interband transition onset at $`4.0`$ eV. As expected, with increasing size the peak width is reduced due to a weaker Landau damping of SP. In order to illustrate the role of interband screening, we also show the results of calculations with $`\mathrm{\Delta }=0`$. Note that decreasing the surface layer thickness by 1.0 a.u. only somewhat reduces the underscreening; the main effect still comes from the larger spatial extent (about 2 a.u.) of sp-band electron spillover. The effect of underscreened surface region on the absorption is two-fold. The large contrast ratio of $`ϵ_d`$ in the bulk and surface regions \[$`ϵ_d(\omega _{sp})5`$\] leads to a lower, as compared to bulk, average value of interband dielectric function in the relevant frequency region. As a result, the peak position for $`\mathrm{\Delta }=1.0`$ a.u. is slightly blueshifted with respect to that for $`\mathrm{\Delta }=0`$ ($`\delta \omega _{sp}0.05`$ eV). At the same time, for $`\mathrm{\Delta }=1.0`$ a.u., the peak amplitude is larger although the resonance width stays unchanged. The stronger absorption for $`\mathrm{\Delta }=1.0`$ a.u. is caused by a weaker screening of the SP electric field, that determines the peak oscillator strength, in the surface region. The calculated local field, $`E`$, at resonance frequency is plotted in Fig. 4 vs. molecule-nanoparticle distance, $`d=r_0R`$. The gradual rise of field magnitude on the length scale of electron spillover replaces the discontinuity (for $`ϵ_d,ϵ_m1`$) of classical field at the sharp boundary. It can be seen that while at $`\mathrm{\Delta }=1.0`$ a.u. the field amplitude is larger than for $`\mathrm{\Delta }=0`$, the difference decreases for larger $`d`$. Correspondingly, the effect of interband screening on Raman signal enhancement, $`|M|^2|E|^4`$, is substantial only when the molecule is located close to the nanoparticle. In Fig. 5, we plot the enhancement factor $`|M|^2`$ as a function of incident light energy for different nanoparticle sizes. Note that the asymmetric shape of the enhancement peak, as compared to the absorption peak, is because the former is determined by the absolute value of polarization, rather than its imaginary part. The general tendency is a decrease of SERS for smaller nanoparticles with the enhancement factor reaching only $`|M|^2100`$ for the smallest nanoparticle size, $`D1.4`$ nm. This decrease is mainly related to the Landau damping of SP: at resonance energy, we have $`E1/\gamma R/v_F`$ yielding $`|M|^2R^4`$. The comparison of results for $`\mathrm{\Delta }=0`$ $`\mathrm{\Delta }=1.0`$ a.u. shown in Figs. 5(a) and 5(b), respectively, shows that reducing the thickness of surface layer even by 1.0 a.u. substantially affects the enhancement. The role of interband screening is most visible in the dependence of SERS on nanoparticle size plotted in Fig. 6 at resonance frequency for nanoparticle diameters in the range 1.4–4.5 nm. With decreasing size, the enhancement factor drops by an order of magnitude, while overall enhancement is reduced for $`\mathrm{\Delta }=0`$. Importantly, the latter effect is more pronounced for smaller nanoparticles; for $`\mathrm{\Delta }=1.0`$ a.u., the enhancement factor is larger by 25% for $`N=3028`$ but by 35% for $`N=92`$ than for $`\mathrm{\Delta }=0`$ indicating a more important role of screening in smaller nanoparticles due to larger surface-to-volume ratio. Thus, the proper account of screening gives a slower (as compared to semiclassical models) decrease of the enhancement as nanoparticles become smaller. Note, finally, that by keeping $`\mathrm{\Delta }`$ constant for different nanoparticle sizes, we somewhat underestimated the screening contribution to SERS; indeed, for smaller nanoparticles, underscreened layer is thicker due to stronger deviations of the confining potential from rectangular shape. ## V Conclusions We developed a microscopic model for surface-enhanced Raman scattering in noble-metal nanoparticles. Our approach incorporates, in a unified manner, all relevant quantum-size effects that determine the magnitude of Raman signal enhancement for nanometer-sized particles. While the Landau damping of surface plasmons leads to a general decrease of the enhancement for small particles, this trend is partially offset by the reduction of interband screening in the metal boundary region. The additional enhancement of local field is substantial only in a close proximity to the metal surface, and it is more pronounced for smaller nanoparticles where the electron density profile deviates strongly from the classical shape As a final remark, we considered here a nonresonant Raman scattering, i.e. the case when the excitation energy of a molecule is much larger than the SP energy $`\omega _{sp}`$. In this case, due to small single-molecule polarizability, only linear (in molecule) response needs to be considered and SERS is dominated by a single back and forth energy transfer within molecule-nanoparticle system \[see Fig. 1(d)\]. In the resonant case, multiple energy transfer processes give rise to a nonradiative width of the molecular levels and, in general, to a reduction of the enhancement.kall-prl04 For small molecule-surface separations, these nonradiative processes are dominated by surface-enhanced electron-hole pair generation in the metal. Even in simple metals, these processes are enhanced due to a reduced intraband (Thomas-Fermi) screening near the boundary. persson-prb82 ; stockman-prb04 The issue of nonradiative decay for noble-metal nanoparticles, where interband screening effects become important, will be addressed in a future publication. ###### Acknowledgements. This work was supported by NSF under Grants No. DMR-0305557 and NUE-0407108, by NIH under Grant No. 5 SO6 GM008047-31, and by ARL under Grant No. DAAD19-01-2-0014. * ## Appendix A Derivation of local potential The full self-consistent potential $`\mathrm{\Phi }(𝐫)=\varphi (𝐫)+\varphi _0(𝐫)`$ sutisfies Poisson equation $`\mathrm{\Phi }(\omega ,𝐫)=\varphi _0(𝐫)+e^2{\displaystyle d^3r^{}\frac{\delta n(\omega ,𝐫^{})}{|𝐫𝐫^{}|}},`$ (16) where the induced density si comprised of sp-band, d-band and medium contriburions, $`\delta n(𝐫)=\delta n_s(𝐫)+\delta n_d(𝐫)+\delta n_m(𝐫)`$. Using Eq. (III) for $`\delta n_m`$ and $`\delta n_d`$ and integrating by parts, Eq. (16) takes the form $`ϵ(r)\mathrm{\Phi }(𝐫)=\varphi _0(𝐫)+e^2{\displaystyle d^3r^{}\frac{\delta n_s(𝐫^{})}{|𝐫𝐫^{}|}}`$ $`+{\displaystyle \frac{ϵ_d1}{4\pi }}{\displaystyle d^3r^{}^{}\frac{1}{|𝐫𝐫^{}|}^{}\theta (R_dr)\mathrm{\Phi }(𝐫^{})}`$ $`+{\displaystyle \frac{ϵ_m1}{4\pi }}{\displaystyle d^3r^{}^{}\frac{1}{|𝐫𝐫^{}|}^{}\theta (rR)\mathrm{\Phi }(𝐫^{})},`$ (17) where $`ϵ(r)=(ϵ_d`$, 1, $`ϵ_m`$) for $`r`$ in the intervals \[$`(0,R)`$, $`(R_d,R)`$, $`(R,\mathrm{})`$\], respectively. Since the source term has the form $`\varphi _0(𝐫)=\varphi _0(r)\mathrm{cos}\theta =eE_ir\mathrm{cos}\theta `$, we expand $`\mathrm{\Phi }`$ and $`\delta n_s`$ in terms of spherical harmonics and, keeping only the dipole term ($`L=1`$), obtain $`ϵ(r)\mathrm{\Phi }(r)=\varphi _0(r)+e^2{\displaystyle 𝑑r^{}r^2u(r,r^{})\delta u_s(r^{})}`$ $`{\displaystyle \frac{ϵ_d1}{4\pi }}R_d^2{\displaystyle \frac{u(r,R_d)}{R_d}}\mathrm{\Phi }(R_d)`$ $`+{\displaystyle \frac{ϵ_m1}{4\pi }}R^2{\displaystyle \frac{u(r,R)}{R}}\mathrm{\Phi }(R),`$ (18) where $$u(r,r^{})=\frac{4\pi }{3}\left[\frac{r^{}}{r^2}\theta (rr^{})+\frac{r}{r^2}\theta (r^{}r)\right]$$ (19) is the dipole term of the radial component of the Coulomb potential. The above equation can be simplified to $`ϵ(r)\mathrm{\Phi }(r)=\overline{\varphi }(r){\displaystyle \frac{ϵ_d1}{3}}\beta (r/R_d)\mathrm{\Phi }(R_d)`$ $`+{\displaystyle \frac{ϵ_m1}{3}}\beta (r/R)\mathrm{\Phi }(R),`$ (20) where $`\beta (r/R)=\left(3R^2/4\pi \right)\left[u(r,R)/R\right]`$ is given by $$\beta (x)=x^2\theta (x1)2x\theta (1x),$$ (21) and we introduced a shorthand notation $`\overline{\varphi }(r)=\varphi _0(r)+e^2{\displaystyle 𝑑r^{}r^2u(r,r^{})\delta n_s(r^{})}.`$ (22) The boundary values of $`\mathrm{\Phi }`$ can be obtained by matching $`\mathrm{\Phi }(r)`$ at $`r=R_d,R`$, yielding $`(ϵ_d+2)\mathrm{\Phi }(R_d)+2a(ϵ_m1)\mathrm{\Phi }(R)=3\overline{\varphi }(R_d),`$ $`(ϵ_d1)a^2\mathrm{\Phi }(R_d)+(2ϵ_m+1)\mathrm{\Phi }(R)=3\overline{\varphi }(R),`$ (23) where $`a=R_d/R`$. Substituting $`\mathrm{\Phi }(R_d)`$ and $`\mathrm{\Phi }(R)`$ back into Eq. (A), we arrive at $`ϵ(r)\mathrm{\Phi }(r)=\overline{\varphi }(r)\beta (r/R_d){\displaystyle \frac{\lambda _d}{\eta }}\left[\overline{\varphi }(R_d)2a\lambda _m\overline{\varphi }(R)\right]`$ $`+\beta (r/R){\displaystyle \frac{\lambda _m}{\eta }}\left[\overline{\varphi }(R)a^2\lambda _d\overline{\varphi }(R_d)\right],`$ (24) where the cofficients $`\lambda `$ are given by Eq. (11). Separating out $`\delta n_s`$-dependent contribution, we arrive at Eq. (8). The expression for nanoparticle polarizability, $`\alpha _p(\omega )`$, can be obtained from the large-$`r`$ asymptotics of induced potential, Eq. (14). From Eqs. (III)–(III), we find for $`rR`$ $$\delta \varphi _d(r)=\frac{eE_i}{r^2}\alpha _d,\delta \varphi _s(r)=\frac{eE_i}{r^2}\alpha _s,$$ (25) with $`\alpha _d=R^3{\displaystyle \frac{(1\lambda _m)a^3\lambda _d\lambda _m}{12a^3\lambda _d\lambda _m}},`$ $`\alpha _s={\displaystyle \frac{4\pi }{3eE_i}}[{\displaystyle _0^{\mathrm{}}}dr^{}r^3\delta n_s(r^{})`$ $`{\displaystyle \frac{\lambda _d\lambda _m+\lambda _d(12a^3)}{12a^3\lambda _d\lambda _m}}{\displaystyle _0^R}𝑑r^{}r^3\delta n_s(r^{})`$ $`{\displaystyle \frac{(1\lambda _m)a^3\lambda _d\lambda _m}{12a^3\lambda _d\lambda _m}}R^3{\displaystyle _R^{\mathrm{}}}𝑑r^{}\delta n_s(r^{})`$ $`+{\displaystyle \frac{(1+\lambda _m)\lambda _d}{12a^3\lambda _d\lambda _m}}{\displaystyle _{R_d}^R}dr^{}(r^3R_d^3)\delta n_s(r^{})].`$ (26) The nanoparticle polarizability is $`\alpha _p=\alpha _d+\alpha _s`$. Note that for frequencies below interband absorption onset, the d-band/medium contribution $`\alpha _d`$ is real.
warning/0506/cond-mat0506430.html
ar5iv
text
# The implications of resonant x-ray scattering data on the physics of the insulating phase of V2O3. ## I Introduction In the last fifteen years, after the discovery of high-temperature superconductivity in cuprates, there has been an upsurge of renewed interest in the electronic properties of transition metal oxides. Among the different techniques applied to investigate such materials, resonant x-ray scattering (RXS) has proved to be a powerful tool to extract direct informations about magnetic and electronic distributions: no other methods have the same flexibility to detect lattice, orbital and magnetic anisotropies within the same experimental setup. Yet, a general theoretical comprehension of some important implications of this technique is still lacking, and this circumstance has sometimes led to incorrect conclusions about the origin of the anomalous signals. This is particularly true for magnetic RXS, where the absence of numerical ab-initio simulations has strongly limited the possibility of quantitative investigations. Only recently few papers have appeared dealing with this matter. In particular Takahashi et al. have used the ab initio local density approximation (LDA) + $`U`$ scheme, taking also into account spin-orbit interaction, to describe such phenomena. In this way they were able to calculate the magnetic RXS in the energy region of the $`4p`$ conduction band in KCuF<sub>3</sub>. One of the purposes of this paper is to present an alternative method to deal with magnetic phenomena, which we believe physically more intuitive, based on a relativistic extension of the Schrödinger Equation and the multiple scattering theory (MST). This extension is obtained by eliminating the small component of the relativistic wave-function in the Dirac Equation and working only with the upper component, thereby using the more intuitive non relativistic set of quantum numbers $`(l,m,\sigma )`$ for the angular and spin momenta. In order to take into account spin-polarized potentials and spin-orbit interaction in the framework of MST we solve a two-channel problem, in each atomic sphere, for the two spin components of the wave-function coupled by the spin-orbit interaction. The condition $`m+\sigma =m^{}+\sigma ^{}`$ must be satisfied, due to the local conservation of the $`z`$-projection of total angular momentum $`\stackrel{}{j}=\stackrel{}{l}+\stackrel{}{\sigma }`$. We have implemented the computer code to calculate magnetic RXS into the FDMNES package. Secondly and more important, we present here the first application of this method to deal with the case of V<sub>2</sub>O<sub>3</sub>, which has attracted considerable attention in the past four years, due to the peculiar interplay of magnetic and orbital ordering. By performing a quantitative analysis of the recent RXS experiments carried out by Paolasini et al. at the vanadium K edge, we shall try to establish the magnetic space group of the monoclinic phase, in the endeavor to resolve some controversies that are still going on in the scientific community. A brief history of the most recent findings about V<sub>2</sub>O<sub>3</sub> will motivate our work. This compound is a Mott-Hubbard system showing a metal-insulator transition at around 150 K from a paramagnetic metallic (PM) to an anti-ferromagnetic insulating (AFI) phase due to the interplay between band formation and electron Coulomb correlation. A structural phase transition takes place, at the same temperature, from corundum to monoclinic crystal class. In recent years, on the basis of an old theoretical model by Castellani et al., Fabrizio et al. suggested that a direct observation of orbital ordering (OO) could have been possible by means of RXS. Soon after Paolasini et al. interpreted the forbidden (111) monoclinic Bragg-forbidden reflection as an evidence of such OO. All this gave rise to an intense debate whose main achievement was to prove the incorrectness of the old Castellani et al. model. In fact, non-resonant x-ray magnetic scattering experiments and absorption linear dichroism showed that the atomic spin on vanadium ions is S=1, and not S=1/2, as supposed in Ref. \[\]. Yet, none of the attempts to explain the origin of the (111) monoclinic reflection can be considered satisfactory, for reasons that will be clearer in the following. Here we just recall that many different physical mechanisms have been proposed, from the antiferro-quadrupolar ordering of the 3$`d`$ orbitals, to the combined action of the ordered orbital and magnetic degrees of freedom; from an orbital ordering associated with a reduction of the magnetic symmetry, to the anisotropies in the magnetic octupolar or toroidal distribution. In this paper we shall focus on a series of Bragg-forbidden reflections measured by Paolasini and collaborators. Of particular interest, from a theoretical point of view, is the set of data of Ref. \[\], where the different reflections have been collected with particular care on their relative intensities, and energy and azimuthal scans have been performed in both the rotated ($`\sigma \pi `$) and unrotated ($`\sigma \sigma `$) channels. Thanks to the wide information content of these data, we are able to demonstrate that the signal cannot be due to any kind of orbital ordering or charge anisotropy and it must be magnetic. Moreover we can show that both mechanisms suggested in Refs. \[\] can be at work for the (1,1,1) reflection, as there is no extinction rule for either dipole-quadrupole or quadrupole-quadrupole transitions. Which of the two contributions dominates depends strongly on the kind of reflection and on the azimuthal angle. To reach this conclusion ab initio calculations have proved necessary. The next section is devoted to the description of our method to deal with magnetic phenomena without resorting to the Dirac equation. We shall work in the framework of MST with muffin-tin approximation or within the finite difference method (FDM), ie, without approximation on the geometrical shape of the potential. In Sec. III we introduce the crystal and electronic properties of V<sub>2</sub>O<sub>3</sub> and compare our results with the experimental data, discussing the physical mechanism behind each reflection. Finally we draw some conclusions on the possibility of gaining information about orbital and magnetic degrees of freedom by means of RXS. ## II Method of calculation ### A RXS and related spectroscopies Core resonant spectroscopies are described by the virtual processes that promote a core electron to some empty energy levels. They all depend on the transition matrix elements of the operator $`\widehat{O}`$ expressing the interaction of electromagnetic radiation with matter: $$M_{ng}=\psi _n|\widehat{O}|\psi _g$$ (1) Here $`\psi _g`$ is the ground state and $`\psi _n`$ the photo-excited state. In the x-ray regime, the operator $`\widehat{O}`$ is usually written through the multipolar expansion of the photon field up to the electric quadrupole term: $$\widehat{O}^{i(o)}=\stackrel{}{ϵ}^{i(o)}\stackrel{}{r}\left(1\frac{1}{2}i\stackrel{}{k}^{i(o)}\stackrel{}{r}\right)$$ (2) where $`\stackrel{}{r}`$ is the electron position measured from the absorbing ion, $`\stackrel{}{ϵ}^{i(o)}`$ is the polarization of the incoming (outgoing) photon and $`\stackrel{}{k}^{i(o)}`$ its corresponding wave vector. In the following dipole and quadrupole will refer to electric dipole and electric quadrupole. In RXS the global process of photon absorption, virtual excitation of the photoelectron, and subsequent decay with re-emission of a photon, is coherent throughout the crystal. Such a coherence gives rise to the usual Bragg diffraction condition, that can be expressed, at resonance, as: $$F=\underset{a}{}e^{i\stackrel{}{Q}\stackrel{}{R}_a}(f_{0a}+f_a^{}+if_a^{\prime \prime })$$ (3) where $`\stackrel{}{R}_a`$ stands for the position of the scattering ion $`a`$, $`\stackrel{}{Q}`$ is the diffraction vector and $`f_0`$ is the usual Thomson factor. The resonant part, $`f^{}+if^{\prime \prime }`$, is the anomalous atomic scattering factor (ASF), given by the expression: $$f^{}+if^{\prime \prime }=\frac{m_e}{\mathrm{}^2}\frac{1}{\mathrm{}\omega }\underset{n}{}\frac{(E_nE_g)^3M_{ng}^oM_{ng}^i}{\mathrm{}\omega (E_nE_g)i\frac{\mathrm{\Gamma }_n}{2}}$$ (4) Here $`\mathrm{}\omega `$ is the photon energy, $`m_e`$ the electron mass, $`E_g`$ the ground state energy, and $`E_n`$ and $`\mathrm{\Gamma }_n`$ are the energy and inverse lifetime of the excited states. In practice, the intermediate states $`\psi _n`$ are in the continuum, normalized to one state per Rydberg. For this reason it is useful to label them by their energy $`E`$ and re-express Eq. (4) as: $`f^{}+if^{\prime \prime }=m_e\omega ^2{\displaystyle \underset{n}{}}{\displaystyle _{E_F}^{\mathrm{}}}{\displaystyle \frac{M_{ng}^oM_{ng}^i}{EE_g\mathrm{}\omega i\frac{\mathrm{\Gamma }(E)}{2}}}𝑑E`$ (5) using the fact that, at resonance, $`E_nE_g\mathrm{}\omega `$. $`E_F`$ is the Fermi energy. In Eq. (5) the summation over $`n`$ is now limited to states having the energy $`E`$. It is now straightforward to make the connection with another spectroscopic technique: x-ray near edge absorption (XANES). In this case the cross-section simply corresponds to the imaginary part of the ASF when $`\widehat{O}_o=\widehat{O}_i`$ (forward scattering). Choosing prefactors in order to have $`f^{\prime \prime }`$ in units of the classical electron radius, $`r_02.8210^{15}`$ m, and the absorption cross section $`\sigma `$ in megabarn, we have the relation: $$\sigma =4\pi 10^{22}\frac{a_0^2\alpha ^3m_ec^2}{\mathrm{}\omega }f^{\prime \prime }$$ (6) where $`a_0`$ is the Bohr radius, $`\alpha `$ is the fine structure constant and $`c`$ is the speed of light. The close connection between the two spectroscopies is thus evident. The use of RXS can have some advantages in exploring electronic properties compared to the absorption techniques. First, with RXS it is possible to select different relative conditions in incoming and outgoing polarizations and wave vectors, and this gives more opportunities to probe the magnetic and electronic anisotropies of the material. Secondly, RXS can be more site selective than XANES, due to the Bragg factor (3). For instance, it is possible, in magnetite, to be sensitive only to octahedral Fe<sup>3+</sup>-sites and not to tetrahedral Fe<sup>3.5+</sup>-sites or, in manganites, to probe only Mn<sup>3+</sup>-ions and not Mn<sup>4+</sup>-ions. This can be achieved at reflections that are Bragg-forbidden off-resonance, but become detectable, in resonant conditions, because of non-symmorphic symmetry elements (glide planes, screw axes) in the crystal space group. In these cases the Thomson factors $`f_0`$ (Eq. 3) drop out and one is just sensitive to the changes of the ASF (Eq. (4)) that depend on the relative electronic and magnetic anisotropies on the ions related by such non-symmorphic symmetry elements. It is indeed in such kinds of reflections that orbital ordering or magnetic scattering were reported and they will be the main subject of the present study. ### B Cartesian Tensor approach In Sec. III we shall deal with the symmetry operations of the crystal space group of V<sub>2</sub>O<sub>3</sub>, in order to evaluate the structure factor for RXS. Because of this, it is very useful to write explicitly the dipole and quadrupole components of the matrix elements in Eq. (5). Remembering that incoming and outgoing x-rays can have different polarizations and wave vectors, we get: $`{\displaystyle \underset{n}{}}M_{ng}^oM_{ng}^i`$ $`=`$ $`{\displaystyle \underset{\alpha \beta }{}}ϵ_\alpha ^oϵ_\beta ^iD_{\alpha \beta }`$ (7) $``$ $`{\displaystyle \frac{i}{2}}{\displaystyle \underset{\alpha \beta \gamma }{}}ϵ_\alpha ^oϵ_\beta ^i(k_\gamma ^iI_{\alpha \beta \gamma }k_\gamma ^oI_{\beta \alpha \gamma }^{})`$ (8) $`+`$ $`{\displaystyle \frac{1}{4}}{\displaystyle \underset{\alpha \beta \gamma \delta }{}}ϵ_\alpha ^oϵ_\beta ^ik_\gamma ^ok_\delta ^iQ_{\alpha \beta \gamma \delta }`$ (9) where $`\alpha `$, $`\beta `$, $`\gamma `$ and $`\delta `$ are cartesian coordinate labels and $`D_{\alpha \beta }`$, $`I_{\alpha \beta \gamma }`$ and $`Q_{\alpha \beta \gamma \delta }`$, the dipole-dipole (dd), dipole-quadrupole (dq) and quadrupole-quadrupole (qq) contributions, respectively. Their explicit expression is given by: $`D_{\alpha \beta }`$ $`=`$ $`{\displaystyle \underset{n}{}}\psi _g|r_\alpha |\psi _n\psi _n|r_\beta |\psi _g`$ (10) $`I_{\alpha \beta \gamma }`$ $`=`$ $`{\displaystyle \underset{n}{}}\psi _g|r_\alpha |\psi _n\psi _n|r_\beta r_\gamma |\psi _g`$ (11) $`Q_{\alpha \beta \gamma \delta }`$ $`=`$ $`{\displaystyle \underset{n}{}}\psi _g|r_\alpha r_\beta |\psi _n\psi _n|r_\gamma r_\delta |\psi _g`$ (12) Note that at a K-edge, in absence of spin-orbit interaction and time-reversal breaking potentials, $`D`$, $`I`$ and $`Q`$ are all real. ### C Calculation of the excited states The main difficulty in the ab-initio evaluation of the ASF is the determination of the excited states. Two different procedures, MST and FDM, have already been developed for non-magnetic cases, and the corresponding packages used in some XANES and RXS applications. In the following we introduce a relativistic extension of the ab initio cluster approach, for both MST and FDM, that is still mono-electronic but includes the spin-orbit interaction and allows to handle magnetic processes. The importance of the FDM procedure to avoid the muffin-tin approximation has already been reported and will not be detailed here again. Note only that with FDM the intermediate states are calculated by solving the Schrödinger equation without any approximation on the geometrical shape of the potential, ie, we do not need to use, as usually in MST approaches, a spherically averaged potential in the atomic spheres and a constant among them. This point is essential to study orbitally ordered materials, where an anisotropic orbital distribution is the key-feature of the system and no spherically averaged potentials would be adequate for its description. Unfortunately, large cluster calculations with FDM are prohibitively long in computing time and require a lot of computer memory: that’s why, when appropriate, MST is preferred. In order to treat magnetic effects in MST we have considered the relativistic extension of Schrödinger equation obtained by solving Dirac equation exactly for the upper component of the wave-function, as described by Wood and Boring. This approach ensures that the singularity of the potential at the nuclear center is of centrifugal type ($`r^2`$) even in the presence of spin-orbit terms (so that the usual fuchsian method of solution can be used). It also allows us to work with the usual non relativistic set of quantum numbers ($`l,m,\sigma `$) of angular and spin momenta, which we find much more intuitive than the relativistic set. Our one particle basis will therefore comprise spin-orbit coupled core states of the form: $`\mathrm{\Phi }_{j,j_z;l_c}^c(\stackrel{}{r})`$ $`=`$ $`R_j^c(r)|j,j_z;l_c)`$ (13) $`=`$ $`R_j^c(r){\displaystyle \underset{m_c\sigma }{}}Y_{l_cm_c}\chi _\sigma \left(l_cm_c1/2\sigma |j,j_z\right)`$ (14) and valence excited states which are multiple scattering solutions of the Schrödinger equation with spin-polarized potentials. In atomic units: $`\{^2+k_e^2V_0(\stackrel{}{r})V_1(\stackrel{}{r})s_z2V_2(\stackrel{}{r})\stackrel{}{\mathrm{}}\stackrel{}{s}\}\mathrm{\Psi }_{\stackrel{}{k}_e,s}(\stackrel{}{r})=0`$ (15) These latters are supplemented by the scattered wave boundary conditions: $`\mathrm{\Psi }_{\stackrel{}{k}_e,s}(\stackrel{}{r})=e^{i\stackrel{}{k}_e\stackrel{}{r}}\chi _sf(\stackrel{}{k}_e,s;\stackrel{}{r},s^{}){\displaystyle \frac{e^{k_er}}{r}}\chi _s^{}`$ (16) In Eq. (13) $`\left(l_cm_c1/2\sigma |j,j_z\right)`$ are the Clebsch-Gordan coefficients. In Eq. (15) $`V_0(\stackrel{}{r})`$ is the average of spin up and spin down potentials, $`V_1(\stackrel{}{r})`$ their difference and $`V_2(\stackrel{}{r})`$ the usual spin-orbit term. Such potentials already embody the relativistic corrections due to the reduction of the upper Dirac component of the wave function to an effective Schrödinger equation. In the muffin-tin approximation the solution inside the *i-th* atomic muffin-tin sphere can be written as: $`\mathrm{\Psi }_{\stackrel{}{k}_e,s}(\stackrel{}{r}_i)={\displaystyle \underset{l}{}}{\displaystyle \underset{m\sigma }{}}{\displaystyle \underset{m^{}\sigma ^{}}{}}R_{lm\sigma }^{lm^{}\sigma ^{}}(r_i)B_{lm^{}\sigma ^{}}^i(\stackrel{}{k}_e,s)Y_{lm}(\widehat{r}_i)\chi _\sigma `$ (17) where at the muffin-tin radius $`\rho _i`$ the radial functions $`R_{lm\sigma }^{lm^{}\sigma ^{}}(r_i)`$ match smoothly to the following combination of Bessel and Hankel functions via the atomic $`t_{lm\sigma ,lm^{}\sigma ^{}}`$ matrices (defined below): $`R_{lm\sigma }^{lm^{}\sigma ^{}}(\rho _i)=j_l(k\rho _i)t_{lm\sigma ,lm^{}\sigma ^{}}^1i\delta _{lm\sigma ,lm^{}\sigma ^{}}h_l^+(k\rho _i)`$ (18) With a proper normalization of these radial functions to one state per Rydberg, the scattering amplitudes $`B_{L\sigma }^i(\stackrel{}{k}_e,s)`$ obey the equations (writing $`L`$ for $`lm`$): $`{\displaystyle \underset{j}{}}{\displaystyle \underset{L^{}\sigma ^{}}{}}M_{L\sigma ,L^{}\sigma ^{}}^{ij}B_{L^{}\sigma ^{}}^j(\stackrel{}{k}_e,s)=\delta _{s\sigma }i^lY_L(\widehat{k}_e)e^{i\stackrel{}{k}_e\stackrel{}{R}_i}`$ where $$M_{L\sigma ,L^{}\sigma ^{}}^{ij}=\left(t_{lm\sigma ,l^{}m^{}\sigma ^{}}^1\delta _{ij}\delta _{ll^{}}\delta _{m+\sigma ,m^{}+\sigma ^{}}+G_{L,L^{}}^{ij}\delta _{\sigma ,\sigma ^{}}\right)$$ is the usual multiple scattering matrix, generalized to spin variables, and $`\stackrel{}{R}_i`$ denotes the position of the *i-th* atom in the cluster with respect to the origin of the coordinates. The atomic $`t`$-matrix $`t_{lm\sigma ,lm^{}\sigma ^{}}`$ describes the scattering amplitude of an electron impinging the atomic potential with angular momentum $`l`$, azimuthal component $`m`$ and spin $`\sigma `$ into a state with quantum numbers $`l,m^{},\sigma ^{}`$. The conservation of the total angular momentum $`\stackrel{}{j}=\stackrel{}{l}+\stackrel{}{s}`$ and its $`z`$-projection implies the constraint $`m+\sigma =m^{}+\sigma ^{}`$. Note that also $`l`$ is unchanged in the scattering process, since in the muffin-tin approximation the potential has spherical symmetry. By introducing as usual the scattering path operator $`\tau _{L\sigma ,L^{}\sigma ^{}}^{ij}`$ as the inverse of $`M_{L\sigma ,L^{}\sigma ^{}}^{ij}`$, the solution for the scattering amplitudes $`B_{L\sigma }^i`$ is given by: $`B_{L\sigma }^i(\stackrel{}{k}_e,s)=\delta _{\sigma s}{\displaystyle \underset{j}{}}{\displaystyle \underset{L^{}}{}}\tau _{L\sigma ,L^{}s}^{ij}i^l^{}Y_L^{}(\widehat{k}_e)e^{i\stackrel{}{k}_e\stackrel{}{R}_j}`$ (19) The scattering path operator $`\tau _{L\sigma ,L^{}\sigma ^{}}^{ij}`$ represents the probability amplitude for the excited photoelectron to propagate from site $`i`$, with angular momentum $`L`$ and spin $`\sigma `$, to site $`j`$ with angular momentum $`L^{}`$ and spin $`\sigma ^{}`$. It is the obvious generalization of the corresponding spin-independent quantity. We now have all the ingredients to perform the intermediate sum in Eq. (5). Using the expression given Eq. (17) for the state $`|\psi _n`$ in the matrix elements of the numerator, the sum over $`n`$ becomes an integral over the escape direction of the photoelectron, $`\widehat{k}_e`$, at fixed energy, $`E=k_e^2`$. When we perform the integral over the energy, we get the product of two smoothly varying radial matrix elements of the type $`R_{lm\sigma }^{lm^{}\sigma ^{}}(r)|r^\nu |R_j^c(r)`$ times the appropriate angular (Gaunt) coefficient. The exponent $`\nu `$ is determined by the polar order of the transition. This is further multiplied by the angular integral of the product of the two scattering amplitudes $`B_{L\sigma }^i`$, $`B_{L^{}\sigma ^{}}^j`$ centered on the absorbing sites. This latter can be simplified by the use of a generalized optical theorem: $`{\displaystyle \underset{s}{}}{\displaystyle 𝑑\widehat{k}_eB_{L\sigma }^i(\stackrel{}{k}_e,s)B_{L^{}\sigma ^{}}^j(\stackrel{}{k}_e,s)}=\mathrm{}\tau _{L\sigma ,L^{}\sigma ^{}}^{ij}`$ (20) As a consequence, the knowledge of $`\tau `$ at all relevant energies, and of the radial solutions $`R_{lm\sigma }^{lm^{}\sigma ^{}}(r)`$ on the photoabsorber are sufficient to calculate the ASF. In the case of V<sub>2</sub>O<sub>3</sub>, in order to construct the spin-polarized atomic potentials, we have used the prescription by von Barth and Hedin: they were derived from the non-self-consistent spin-polarized charge density obtained by superimposing the atomic charge densities with two magnetic electrons on each Vanadium ion, as suggested by the experimental data. We believe that this approximation is not far from reality, as shown a posteriori by the goodness of the results. Given the potential, it is straightforward to calculate the atomic spin dependent $`t`$-matrix: we just solve the two-channel problem arising from the Schrödinger equation (15) due to the presence of the spin-orbit potential. In this way this latter is treated on the same footing as the other potentials and not in perturbation theory as usually done. The relativistic extension in the FDM scheme follows closely the previous treatment for MST. The space is partioned, as in the multiple scattering approach, in three regions: i) An outer sphere surrounding the cluster where one impose the scattering behaviour of the wave function given in Eq. (16). ii) An atomic region made up of spheres, around each atom, with radii much smaller than the muffin-tin radii (of the order of 1 a.u. or less). Here the charge density is to a very good approximation spherically symmetric, due to the presence of the core electrons. iii) Finally, an interstitial region where the Laplacian operator of the Schrödinger equation (15) is discretized and the solution is generated on a greed without any approximation on the geometrical shape of the potential. By imposing a smooth continuity of the overall wave function across the boundaries of the three regions, one can determine both the expansion amplitudes of the wave function inside each atom and the scattering T-matrix of the whole cluster in Eq. (16). Note that in this case the spin-orbit potential is not spherically symmetric. We have checked that for a small cluster with close-packed geometry, this approach and the multiple scattering approach with muffin-tin approximation provide almost identical cross sections. Since V<sub>2</sub>O<sub>3</sub> is a close-packed structure, if one neglects vanadium voids, we have used MST in the muffin-tin approximation for the magnetic calculations. The FDM scheme was instead used in the calculation of the ASF in the case of orbital ordering. ## III The case of V<sub>2</sub>O<sub>3</sub> In this section we specialize to the case of V<sub>2</sub>O<sub>3</sub>, to investigate whether the RXS reflections observed by Paolasini and collaborators are of magnetic or orbital nature. Our result is that magnetic ordering can explain the experimental data with a reasonably good agreement (even if not perfect, for reasons that will be discussed below). On the other hand, we are able to demonstrate that any OO origin of the signal has to be excluded, as it would not fit the energy scan at dipolar energies. Next subsection is devoted to the description of the magnetic space group in the AFI phase, as well as the derivation of the anomalous RXS amplitude. In subsection B we analyze the ab-initio calculations for the magnetic RXS, through the procedure previously developed. Finally, in the last part we show that the consequences of a time-reversal breaking OO are not compatible with the experimental data, thus putting a severe constraint on the possible OO theories for V<sub>2</sub>O<sub>3</sub>. ### A Crystal and magnetic structure As a first step we shall try to establish the magnetic symmetry group in the AFI phase. Our starting point is the crystallographic structure, as given by Dernier and Marezio together with the magnetic data given by Moon and later confirmed by Wei Bao et al. Referring to the frame and the numbering of vanadium atoms of Fig. 1, we can divide the eight atoms of the monoclinic unit cell into two groups of four, $`V_1,V_2,V_3,V_4`$ and $`V_1^{}`$, $`V_2^{}`$, $`V_3^{}`$, $`V_4^{}`$, with opposite orientations of the magnetic moment. The two groups are related by a body-centered translation. Vanadium-ion magnetic moments, indicated by arrows in the figure, lie perpendicular to the $`b_m`$-axis at an angle of 138<sup>0</sup> away from the monoclinic $`c_m`$-axis. Neglecting the magnetic moments, the previous two groups of four atoms with their oxygen environments are translationally equivalent. We can infer from these data that the magnetic space group can be written as $`P2/a+\widehat{T}\{\widehat{E}|t_0\}P2/a`$, where $`\widehat{T}`$ is the time-reversal operator and $`t_0`$ the body-centered translation. The monoclinic group $`P2/a`$ contains four symmetry operations: the identity $`\widehat{E}`$, the inversion $`\widehat{I}`$, the two-fold rotation about the monoclinic $`b_m`$ axis $`\widehat{C}_{2b}`$ and the reflection $`\widehat{m}_b`$ with respect to the plane perpendicular to this axis. All these operations are associated to the appropriate translations, as shown in the following table: $`\begin{array}{ccccc}\mathrm{atom}& \mathrm{position}& \mathrm{spin}& \mathrm{sym}.& \mathrm{translation}\\ & & & & \\ V_1& (u,v,w)& & \widehat{E}& 0\\ V_2& (u,v,w)& & \widehat{I}& 0\\ V_3& (\frac{1}{2}+u,v,w)& & \widehat{T}\widehat{m}_b& (\frac{1}{2},0,0)\\ V_4& (\frac{1}{2}u,v,w)& & \widehat{T}\widehat{C}_{2b}& (\frac{1}{2},0,0)\\ V_1^{}& (\frac{1}{2}+u,\frac{1}{2}+v,\frac{1}{2}+w)& & \widehat{T}& (\frac{1}{2},\frac{1}{2},\frac{1}{2})\\ V_2^{}& (\frac{1}{2}u,\frac{1}{2}v,\frac{1}{2}w)& & \widehat{T}\widehat{I}& (\frac{1}{2},\frac{1}{2},\frac{1}{2})\\ V_3^{}& (u,\frac{1}{2}v,\frac{1}{2}+w)& & \widehat{m}_b& (0,\frac{1}{2},\frac{1}{2})\\ V_4^{}& (u,\frac{1}{2}+v,\frac{1}{2}w)& & \widehat{C}_{2b}& (0,\frac{1}{2},\frac{1}{2})\end{array}`$ TABLE I. Here $`u=0.3438,v=0.0008,w=0.2991`$ are the fractional coordinates of the atoms in unit of the monoclinic axis. Notice that it is sufficient to calculate the ASF for atom number 1, since all the others can be deduced by applying the symmetry operation indicated in the fourth column of the table. In the following, we re-analyze the extinction rules for the different reflections $`(h,k,l)`$ measured by Paolasini et al. in terms of the crystal tensors in the dd, dq and qq channels (Eq. (12)). All the recorded reflections have the same incoming polarization, $`\sigma `$ (electric field perpendicular to the diffraction plane), whereas the outgoing polarization is analyzed both in the $`\sigma `$ and $`\pi `$ channels (this latter with the electric field in the diffraction plane). The azimuthal scans are then registered by rotating the sample around the diffraction vector. We refer to Paolasini et al. \[\] for the definition of the azimuthal origin: the situation where the scattering plane contains $`a`$ and $`c`$ hexagonal axes corresponds to an azimuthal angle of 90 degrees. From equation (3) and Table I, neglecting the very small component of $`\stackrel{}{R}`$ along $`b_m`$, ($`v=0.0008`$), we get for each reflection $`(h,k,l)`$ the expression: $`^{hkl}=(1+()^h\widehat{T}\widehat{m_b})(e^{i\varphi _{hkl}}+e^{i\varphi _{hkl}}\widehat{I})`$ (22) $`(1+()^{h+k+l}\widehat{T})f_1`$ (23) where $`R(u,0,w)`$ is the position of the $`V_1`$ atom, $`\varphi _{hkl}\stackrel{}{Q}\stackrel{}{R}2\pi (hu+lw)`$ and $`f_1`$ stands for the ASF as defined in (Eq. 5). Note that $`f_1`$ is a scalar and we adopt the convention, when we say that a symmetry operator acts on $`f_1`$, that it acts only on the tensor components (dd, dq and qq) given by Eq. (12). Focusing on the reflections with $`h+k+l=`$odd and noticing that $`D_{\alpha \beta }`$ and $`Q_{\alpha \beta \gamma \delta }`$ are inversion-even, while $`I_{\alpha \beta \gamma }`$ is inversion-odd, we find that only the following tensor components contribute to the signal: $`𝒟_{\alpha \beta }^{hkl}=4i(1+(1)^{h+n_y+1})\mathrm{cos}(\varphi _{hkl})\mathrm{}(D_{\alpha \beta })`$ (24) $`_{\alpha \beta \gamma }^{hkl}=4(1+(1)^{h+n_y+1})\mathrm{sin}(\varphi _{hkl})\mathrm{}(I_{\alpha \beta \gamma })`$ (25) $`𝒬_{\alpha \beta \gamma \delta }^{hkl}=4i(1+(1)^{h+n_y+1})\mathrm{cos}(\varphi _{hkl})\mathrm{}(Q_{\alpha \beta \gamma \delta })`$ (26) where $`n_y`$ is the number of $`y`$ labels among the tensor indices $`\alpha `$, $`\beta `$, $`\gamma `$, and $`\delta `$. Thus, in order to have a signal, $`h`$ and $`n_y`$ must have different parity. Notice that all the quantities in Eq. (26) are magnetic, as only imaginary parts of cartesian tensors are involved. Moreover, the three scattering amplitudes are all purely imaginary, as the dq polarizations carry an extra imaginary unit (see Eq. 9). As a consequence, they all interfere. A separate analysis of the three tensors of Eq. (26) gives the following indications. In the case of dd tensors, when $`h`$ is odd, the only non-zero component is $`𝒟_{xz}^{hkl}𝒟_{zx}^{hkl}L_y`$. When $`h`$ is even the non-zero components are $`𝒟_{xy}^{hkl}𝒟_{yx}^{hkl}L_z`$ and $`𝒟_{yz}^{hkl}𝒟_{zy}^{hkl}L_x`$. As the magnetic moment direction is perpendicular to the $`b_m`$-axis, $`L_y`$ is zero. Thus for odd $`h`$ no signal is expected in the dipolar region. On the contrary, when $`h`$ is even, a dd contribution is present. Of course, no $`\sigma \sigma `$ magnetic scattering is allowed. These facts explain why the experimental spectrum for the $`(2,\overline{2},1)_{\sigma \pi }`$ shows structures at the $`4p`$ edge, contrary to the $`(2,\overline{2},1)_{\sigma \sigma }`$ reflection and to the $`(1,1,1)`$ and $`(3,\overline{1},1)`$ reflections for both polarization conditions. Notice that these results had been already derived with similar methods. These remarks do not apply for qq and dq tensors. They always contribute, for all the investigated reflections, in both $`\sigma \sigma `$ and $`\sigma \pi `$ channels, except at some very specific azimuthal angles. At the K edge these tensors measure the transitions to the states with pure 3$`d`$ or hybridized $`3d`$-$`4p`$ character. For this reason they have a finite value just close to the Fermi energy. The physical quantities measured by these operators can be identified as the octupolar magnetic moment for the qq term and the toroidal or quadrupolar magnetic moment for the dq tensor. ### B Analysis of the magnetic signal In order to get the absolute intensities and shape of the spectra, we need now to resort to ab initio calculations. We performed such calculations for the crystal and magnetic structures given in Table I. Thus, we do not neglect the small $`v`$ value but we shall see that the conclusion given in the previous subsection is not modified. The potential is calculated using a superposition of atomic densities obtained from an atomic, self-consistent Hartree-Fock calculation, with a $`3d^2`$, spin 1, configuration. The MST approach is used with different cluster radii from 3.0 up to 7.2 Å, ie, from a $`VO_6`$ molecule to a cluster containing 153 atoms. We find that, in order to get the shoulder at 5475 eV in both XANES and $`(2,\overline{2},1)_{\sigma \pi }`$ spectra, as well as the 5487 eV shoulder in the $`(2,\overline{2},1)_{\sigma \pi }`$, we need the biggest cluster radius (7.2 Å - see Fig. 2). Nevertheless the main features are present for all Bragg peaks even for the $`VO_6`$ molecule calculation. The behaviour of the azimuthal scans against the cluster radius looks more complex. The agreement improves up to the 4.3 Å, corresponding to 33 atoms, and then decreases when the cluster radius is further increased. The reason for this is discussed below. For the moment we keep the 4.3 Å radius and compare such energy and azimuthal spectra with the experimental ones, as shown in Figs. 3 and 4. Looking at the spectra shown in Fig. 3, we can claim a satisfactory experiment-theory agreement for the different reflections. In particular we get the very thin resonance in the $`3dt_{2g}`$ energy range at the $`(1,1,1)`$ and $`(3,\overline{1},1)`$ peaks for both $`\sigma \sigma `$ and $`\sigma \pi `$ polarization conditions and for the $`(2,\overline{2},1)_{\sigma \sigma }`$ reflection. The broad intensity in the $`4p`$ energy range in the $`(2,\overline{2},1)_{\sigma \pi }`$ is also reproduced. The intensity ratio between $`\sigma \sigma `$ and $`\sigma \pi `$ channels and between the various reflections are also quite good considering the difficulties on both experimental and theoretical sides, even if for the $`(1,1,1)_{\sigma \sigma }`$ reflection an extra factor improves the agreement with the experiment. Note that, differently from all other reflections, the $`(2,\overline{2},1)_{\sigma \pi }`$ can contain some non-resonant component not taken into account in our calculation. Moreover, a non-zero offset seems to be present at $`(2,\overline{2},1)_{\sigma \sigma }`$ on the experimental side, and this is probably responsible for the discrepancy in the corresponding energy and azimuthal scans. The azimuthal scans, shown in Fig. 4, are reasonable for all $`\sigma \pi `$ reflections. Only the $`(1,1,1)_{\sigma \sigma }`$ is much less satisfying: yet, the corresponding data belong to the first experiment (Ref. , while the others belong to Ref. ), and this does not allow to be sure about the relative intensity as in the other cases. Finally, we are also able to reproduce the $`\pi `$ and $`2\pi `$ periodicity, expected from the low symmetry of the compound, for the $`\sigma \sigma `$ and $`\sigma \pi `$ channels, respectively. We judge the agreement between theory and experiment quite satisfactory. Indeed the interference between the dd, dq and qq parts makes the problem really tricky. Foe example, $`p`$ orbitals are far more delocalized than $`d`$ ones. For this reason they react in a different way to the core hole screening. In a first approximation $`d`$ states are more shifted towards lower energy than $`p`$ states. Thus, the dq terms, which probe the hybridization between $`p`$ and $`d`$ orbitals, are strongly influenced by the details of the electronic structure around the photo-absorbing ion. If we consider that V<sub>2</sub>O<sub>3</sub> is a strongly correlated electron system, with a quantum-entangled ground-state wave function, we find quite surprising that we have reached such a good agreement in the $`3d`$ energy region with a one-electron calculation. Our idea is that this can help understanding which features can be explained in terms of an independent particle approach and which cannot (see below). Note that the localization of the $`d`$ orbitals also explains why the azimuthal scans look better when the calculations are performed using the smaller 4.3 Å cluster radius. To end this discussion we stress that energy spectra profiles around 5464 eV are very sensitive to the azimuthal angle and, vice-versa, azimuthal profiles are very sensitive to the energy. This is due to the fact that qq and dq components have both strong (and different) angular and energy dependence. There are some angles where the calculated profiles appear doubled, with about 1 eV between the two peaks. For instance, this is what happens at the experimental azimuth of the $`(3,\overline{1},1)_{\sigma \sigma }`$ energy spectrum: a simple shoulder appears at 5465 eV, well described by our calculation. Finally, we want to comment about the two papers (by Lovesey et al. and by Tanaka.) that have already pointed out the magnetic origin of the $`h+k+l=`$odd reflections. These works disagree on the physical mechanism of the signal, as for Tanaka only dq terms contribute, while Lovesey and coworkers attribute the signal to a pure qq reflection. Our results on this point show that both dq and qq channels can contribute to the global intensity depending on the reflections (whether $`h=`$odd or $`h=`$even) and on the azimuthal angle (see Fig. 5). At the prepeak energy, E=5465 eV, the dq term represents the strongest contribution for the $`(1,1,1)`$ and $`(3,\overline{1},1)`$ reflections, even if the qq term is not completely negligeable. For the $`(2,\overline{2},1)_{\sigma \sigma }`$, on the contrary, the qq contribution is clearly the dominating one. Finally, it is interesting to note that for the $`(2,\overline{2},1)_{\sigma \pi }`$ reflection the dd term plays the major role even at the $`3d`$ energies. Anyway the qq contribution has always a non-negligeable influence on the total signal. Comparing our findings with those of Refs. we can state that there are no strict extinction rules regarding dq and qq terms for all these reflections. Nonetheless, the main contribution to $`h+k+l=`$odd, $`h=`$odd reflections comes from the dq channel (as in Ref. ), while the main term in $`h+k+l=`$odd, $`h=`$even reflections is of qq origin (as in Ref. ). Such a tensor analysis allows also to identify a feature that is not correclty described by our one-electron calculations, ie, the direction of the magnetic moment. Indeed, a small $`L_y`$ component (about $`10\%`$ the global moment) is found to contribute to the $`h=`$odd signal (see Fig. 5), contrary to what expected by our previous theoretical analysis. This result can be explained by noticing that in order to obtain the correct direction of the magnetic moment a molecular, correlated ground-state wave function is needed, which is far beyond the possibilities of our monoelectronic approach. ### C Analysis of the ”orbital” signal In this subsection, we analyze the effect of a time-reversal breaking OO on the signal, in order to determine whether it can affect our previous results. This is not a secondary issue since the original experimental interpretation suggested that the (111)-monoclinic reflection was due to the OO and a big controversy around this point arose in the literature. The reason why we focus on a ”time-reversal breaking” OO is that, for $`h+k+l=`$odd, the signal is proportional to $`(\widehat{E}\widehat{T})f_1`$, as clear from Eq. (23), so that a non-magnetic signal is allowed only when the time-reversal symmetry is broken. This is obtained, for example, when V<sub>i</sub> and V$`_i^{}`$ have a different orbital occupancy, and we call such a situation time-reversal breaking OO. Different types of orbitally-ordered ground states have been suggested in the literature, but none of them possesses such a feature. As a consequence, non-magnetic signals at the (111)<sub>m</sub> reflection are not expected, unless some ad hoc hypoteses are made, as recognized in the discussion of Sec. VII.D of Ref. \[\]. One possible way to get a signal of OO origin is to consider the first excited state found in Ref. \[\], which has a magnetoelectric (ME) symmetry. It is very close to the ground state ($`1`$ meV) and, because of this, it could be partly occupied. Independently of this, if we consider the two maximal ME subgroups of the full space groups, ie, P2’/a (with symmetry elements $`\widehat{E}`$, $`\widehat{T}\widehat{I}`$, $`\widehat{T}\widehat{C}_{2b}`$, $`\widehat{m}_b`$) and P2/a’ (with symmetry elements $`\widehat{E}`$, $`\widehat{T}\widehat{I}`$, $`\widehat{C}_{2b}`$, $`\widehat{T}\widehat{m}_b`$), it is in principle possible to get a non-magnetic signal at the (111)<sub>m</sub> reflection. In fact, in both cases there are two subgroups of four ions whose electronic densities have anisotropies not connected by time-reversal nor any other symmetry operation. Consider, for example, the case of P2/a’, that could be responsible for non-reciprocal dichroism: the two groups of atoms ($`V_1,V_2^{},V_3,V_4^{}`$) and ($`V_1^{}`$, $`V_2`$, $`V_3^{}`$, $`V_4`$) are independent and the structure factor, Eq. 23, becomes $`^{hkl}=(1+()^h\widehat{T}\widehat{m_b})(e^{i\varphi _{hkl}}+()^{h+k+l}e^{i\varphi _{hkl}}\widehat{T}\widehat{I})`$ (27) $`(f_1+()^{h+k+l}f_1^{})`$ (28) The analogous of Eq. 26 for reflections with $`h+k+l=`$odd is now: $`𝒟_{\alpha \beta }^{hkl}=4i(1+(1)^{h+n_y})\mathrm{sin}(\varphi _{hkl})\mathrm{}(D_{\alpha \beta })`$ (29) $`+4i(1+(1)^{h+n_y+1})\mathrm{cos}(\varphi _{hkl})\mathrm{}(D_{\alpha \beta })`$ (30) $`_{\alpha \beta \gamma }^{hkl}=4(1+(1)^{h+n_y})\mathrm{cos}(\varphi _{hkl})\mathrm{}(I_{\alpha \beta \gamma })`$ (31) $`+4(1+(1)^{h+n_y+1})\mathrm{sin}(\varphi _{hkl})\mathrm{}(I_{\alpha \beta \gamma })`$ (32) $`𝒬_{\alpha \beta \gamma \delta }^{hkl}=4i(1+(1)^{h+n_y})\mathrm{sin}(\varphi _{hkl})\mathrm{}(Q_{\alpha \beta \gamma \delta })`$ (33) $`4i(1+(1)^{h+n_y+1})\mathrm{cos}(\varphi _{hkl})\mathrm{}(Q_{\alpha \beta \gamma \delta })`$ (34) Thus, the breakdown of time-reversal symmetry, with the introduction of the two ASF $`f_1`$ and $`f_1^{}`$, has made possible also non-magnetic scattering, through the real part of the tensor components. Notice that, due to the magnetic space group, both magnetic and non-magnetic scattering can in principle interfere at particular polarization conditions and reflections. In order to calculate the signal explicitly, we need to express the orbital wave functions of V<sub>1</sub> and V$`_1^{}`$ ions. Starting from the orbital occupancy of Refs. , we get, in the reference frame of Fig. 1: $`\varphi _1={\displaystyle \frac{1}{3}}d_{xy}+{\displaystyle \frac{\sqrt{2}}{3}}d_{yz}+{\displaystyle \frac{1}{\sqrt{6}}}d_{z^2}{\displaystyle \frac{1}{3\sqrt{2}}}d_{xz}{\displaystyle \frac{2}{3}}d_{x^2y^2}`$ $`\varphi _1^{}={\displaystyle \frac{2}{3}}d_{xy}+{\displaystyle \frac{1}{3\sqrt{2}}}d_{yz}+{\displaystyle \frac{1}{\sqrt{6}}}d_{z^2}{\displaystyle \frac{\sqrt{2}}{3}}d_{xz}{\displaystyle \frac{1}{3}}d_{x^2y^2}`$ (35) From this orbital occupancy, we can calculate the electron density. Then, solving Poisson’s equation, we get the Coulomb potential. The energy-dependent exchange-correlation potential is then obtained with conventional procedures. The calculation of the ASF is performed for the $`V_1`$ and $`V_1^{}`$ atoms with the FDM option. We kept a rather small cluster radius (3.0 Å), but included the oxygen octahedra as well as the first shell of four vanadium neighbors. The ASF of the other atoms are then deduced by means of the magnetic symmetries of Table I for the two separate subgroups (see also Ref. \[\], sec. VII.D, for a derivation) and the total scattering amplitude is obtained using Eq. (3). Our results for the $`(1,1,1)_m`$ reflection are shown in Fig. 6. As in the experiment, a structure is obtained at the 3d-energy level. Yet, a much bigger signal is present at the 4p edge where the experimental intensity shows no features. Such a disagreement is the proof that the OO cannot be responsible of the signal: The time-reversal breaking OO makes the $`(111)_m`$ reflection allowed not only at the 3d-energy level, but also at the 4p-energies. Note that this result remains valid for any time-reversal breaking OO, not only the one proposed in Eq. (35), as should have been qualitatively expected. Indeed, it has already been shown that the 3d-4p hybridization, even if small, must give a contribution to the dipolar K edge signal. In fact, previous numerical simulations in another compound, LaMnO<sub>3</sub>, have shown that the effect of a symmetry-breaking OO of 3d orbitals produces a signal also at the 4p-energy levels. In this latter case the contribution coming from the Jahn-Teller distortion overwhelms the OO signal. However in the present case, since V<sub>1</sub> and V$`_1^{}`$ have the same local distortion, such a signal should have been detected, as clear from our numerical simulation shown in Fig. 6. Thus, the experimental evidence of absence of any signal at the 4p energies proves two facts: first, there is no time-reversal breaking OO; second, the (111)<sub>m</sub> cannot be due to OO. A direct consequence of the previous analysis is that no magnetoelectric s ubgroups are compatible with the (111)<sub>m</sub> energy scan, as they all break the time-reversal symmetry, in keeping with the negative experimental evidence for magnetoelectricity in V<sub>2</sub>O<sub>3</sub>. This strongly supports the results of Ref. \[\]. Notice that a previous study on the same reflection in V<sub>2</sub>O<sub>3</sub> ended up with the conclusion that the OO was responsible of the signal. Unfortunately that calculation was performed only around the 3d-energy range, where the agreement in the azimuthal scan was quite good and the big feature at the 4p energies was not detected. The lesson to be drawn from this fact is that the azimuthal scan around the diffraction vector is not always a fundamental feature in determining the origin of the reflection. In fact, it usually reflects more the geometry than the electronic properties of the material. This is a rather general comment, not limited to V<sub>2</sub>O<sub>3</sub>: when there are only one or two ab initio independent factors (the tensor components), the angular properties are much more determined by the geometry (ie, the symmetries) than by the dynamics (ie, the relative weight of the radial matrix elements). In this situation only the combined analysis of both energy and azimuthal scans is a reliable tool to investigate the electronic origin of RXS phenomena. ## IV Conclusion The main results of the present paper can be inferred from the conclusions of the last two subsections discussing the results appeared in the literature. In fact, up to now, it was possible to classify the interpretations of the forbidden Bragg reflections with $`h+k+l=`$odd and $`h=`$odd along two main lines of thought: those who explained such reflections in terms of orbital ordering, both associated or not with a reduction of the magnetic symmetry, and those who were inclined to a magnetic origin. A third explanation, that of the antiferroquadrupolar ordering proposed in Ref. \[\] had already been ruled out in Ref. \[\]. Whether the former or the latter interpretations were correct, was not simple to decide on the basis of the azimuthal scans, only. In fact, such scans, for a given multipolar channel, measure the crystal symmetry, rather than the electronic origin of the reflections. The proof of such a statement is that three different mechanisms (magnetic in the qq channel, magnetic in the dq channel and non-magnetic, via orbital ordering) gave all a rather good agreement in the azimuthal scan at the pre-K edge. What definitively rules out the OO origin of such reflections is the energy scan shown in Fig. 6. As noted previously, a direct consequence of this result is that no reduction of the magnetic space group $`P2/a+\widehat{T}\{\widehat{E}|t_0\}P2/a`$ is present. The second important result of the present paper lies in the fact that we were able to make a complete ab-initio analysis of all $`h+k+l=`$odd reflections, starting from the crystal and magnetic structure, only. We showed that their origin (whether dd, dq or qq) strongly depends on the kind of reflection (ie, $`h=`$even or $`h=`$odd), photon energy and azimuthal angle. For $`h=`$odd, the dq channel is predominant, in keeping with the cluster calculations of Tanaka. In spite of this particular agreement, we believe that it is important to stress that our approach goes beyond the cluster calculations performed in Ref. \[\] as well as the simple fitting procedure of Ref. \[\]. In fact only by means of an ab-initio procedure it is possible to cover all the main experimental evidence, ie, the azimuthal behaviour, the energy profiles, and the order of magnitude of the intensity, that is well in keeping with the rough estimate given in Ref. \[\]. We would like to acknowledge L. Paolasini for several useful discussions.
warning/0506/astro-ph0506252.html
ar5iv
text
# Formation and Evolution of Planetary Systems: Upper Limits to the Gas Mass in HD 105 ## 1 Introduction Observations of young stars indicate that circumstellar disks of gas and dust are common in the earliest stages of evolution (e.g., Beckwith & Sargent 1996, Hillenbrand et al. 1998, Haisch et al. 2001, Weinberger et al. 2004, Natta et al. 2004). Most of these studies rely on infrared, submillimeter, or millimeter wavelength emission as the signature of small ($``$ 1 mm) warm dust particles, which trace only a small fraction of the circumstellar gas and solid particles. In their initial active accretion phase, circumstellar disks are composed of primordial molecular cloud material, possibly somewhat processed by the collapse through the accretion shock onto the disk, with gaseous hydrogen and helium constituting $`99\%`$ of the total mass and with $`1\%`$ of the disk mass in small dust particles. As infall from the cloud onto the star+disk system diminishes, the gas and dust components of the disk evolve, presumably losing most of the gas via accretion onto the central star or photoevaporation back to the interstellar medium (Hollenbach et al. 1994, Johnstone et al. 1998, Clarke et al. 2001, Adams et al. 2004). Some of the small dust grains presumably coagulate, growing to eventually form planetesimals and rock/ice planets (e.g., Pollack et al. 1996, Weidenschilling 1977, 1997). During these processes, the circumstellar disk undergoes a transition from being optically thick and (mostly) gaseous to becoming optically thin and (mostly) dusty. There have been many theoretical and observational studies of disks in various stages of evolution pointing towards such a sequence (e.g., Weidenschilling 1997, Suttner & Yorke 2001, Throop et al. 2001, Przygodda et al. 2003, Hogerheijde et al. 2003, Testi et al. 2003, van Boekel et al. 2003, Wolf et al. 2003, Dullemond & Dominik 2004, Kessler-Silacci et al. 2005). Such studies have, however, concentrated only on the more readily observed dust emission from the disk, and suggest the evolution of dust to larger particle sizes and hence lower opacity (Miyake and Nakagawa, 1993). Observations of gas and its evolution are in their infancy. Most studies to date have probed either at near infrared wavelengths (often CO vibrational transitions) the very hot inner surface regions less than about 1 AU from the star (Carr, Mathieu & Najita 2001, Brittain et al. 2003, Najita et al. 2003, Blake & Boogert 2004, Rettig et al. 2004, Thi et al. 2005), or at millimeter wavelengths (often CO rotational transitions) extended gas–rich disks that are hundreds of AU in size (e.g., Skrutskie et al. 1991, Zuckerman et al. 1995, Dutrey et al. 1998, Duvert et al. 2000, Pietu et al. 2003, Qi et al. 2003, , Dent et al. 2005, Greaves 2005). The millimeter observations are generally not sensitive to gas at $``$ 30 AU because the lines become beam–diluted and weak. In addition, the trace species used as millimeter wavelength probes of gas may freeze out onto grain surfaces at $``$ 30 AU, making these observations somewhat unreliable for estimating gas masses (e.g., van Dishoeck 2004). The presence or absence of gas in the planet-forming 1–30 AU regions is difficult to constrain from observations in the near infrared or at millimeter wavelengths. Temperatures in these regions may range from 300 –50 K, and warm gas emitting at these temperatures is best probed in the mid–infrared lines of H<sub>2</sub> at 28.2 and 17.0 $`\mu `$m, and the mid–infrared fine structure lines of abundant ions and atoms, like \[SI\] 25.2 $`\mu `$m (Gorti & Hollenbach 2004, hereafter GH04). These lines are difficult or impossible to observe from the ground. Furthermore, the strong dust continuum of very young objects makes the line to continuum ratio quite low, and the background noise high, unless the lines are spectroscopically resolved (requiring resolving powers from 30,000–100,000). The Infrared Space Observatory (ISO) provided some tantalizing evidence for long–lived gas–rich disks by observing mid–infrared emission lines of H<sub>2</sub> toward debris disk candidates (Thi et al. 2001). However, ground–based observations (Richter et al. 2002, Sheret et al. 2003, Sako et al. 2005), FUSE results (LeCavalier des Etangs et al. 2001), and Spitzer observations (Chen et al 2004) have called some of these results into question. Recently, ground-based IR and space-based UV observations have detected rovibrational (e.g., Bary et al. 2003) and fluorescent (e.g., Herczeg et al. 2002) H<sub>2</sub> emission towards young disks. The fluorescent H<sub>2</sub> emission is found to occur from gas that is warm and probably located within a few AU of the star. The narrow linewidths of the rovibrational IR emission suggests that the emission occurs at much larger distances from the star ($`>`$10 AU), and arises either from non-thermal UV pumping or from the X-ray heating of a very small amount of H<sub>2</sub> gas located in the low-density upper atmosphere. The likely non-thermal excitation of the H<sub>2</sub> seen both in the IR and UV makes it difficult to convert the detected line strengths into gas masses. The nature of gas and dust evolution in circumstellar disks is an intriguing problem for various reasons. A determination of the gas dispersal timescales is important for understanding the formation of planetary systems, and specifically for discriminating between the two competing theories that exist for gas giant planet formation. Core accretion theories for forming gas giants (e.g. Lissauer et al. 1993, Pollack et al. 1996, Kornet et al. 2002, Hubickyj, Bodenheimer, & Lissauer 2004) require long ($`1`$ Myr) gas disk lifetimes to facilitate the formation of rocky cores of a few Earth masses and their subsequent gas accretion to form planets. Detection of gas in older ($`3\times 10^6`$ year) disks would help validate the efficacy of this scenario. The alternate theory of gravitational instability in disks (e.g., Boss 2003) allows the formation of gas giants quickly and the gas may dissipate rapidly. The presence of small amounts of gas in the “terrestrial” zone, $`0.15`$ AU, in later evolutionary stages ($``$ 10 Myrs) influences the dynamics of the smaller planets and planetesimals and is decisive in the final outcome of the planetary system (Agnor & Ward 2002; Kominami & Ida 2002, 2004). Only a narrow range of gas masses ($`0.01`$ M<sub>J</sub>) in this stage of terrestrial planet formation allow for the formation of an Earth-sized planet in an orbit as circular as the Earth. At even later stages ($`>10^7`$ years) in disk evolution, small masses of gas ($`0.01`$ M<sub>J</sub>) can affect dust dynamics via drag forces and hence significantly influence disk structure (Klahr & Lin 2001, Takeuchi & Artymowicz 2001, Takeuchi & Lin 2002), weakening interpretations of gaps and ring signatures as being due to the presence of planets. This paper highlights the potential of the recently launched Spitzer Space Telescope in detecting small amounts of gas (or setting stringent limits on gas mass) in the planet–forming regions of disks around nearby young stars. The Spitzer Space Telescope is able to observe the thermal emission from deep layers in the planet-forming zone, sampling significant mass in this 0.3–30 AU intermediate zone inaccessible to most other techniques. Gas in these regions is likely to be warm ($`100`$ K) and line emission at the mid-infrared ($`1037\mu `$m) wavelengths seen by the Spitzer Infrared Spectrometer (IRS) instrument can probe total gas masses as small as $`0.01`$M<sub>J</sub> in the inner ($`<`$ 20 AU) regions around nearby ($`30`$ pc) stars (GH04). Gas line diagnostics in the Spitzer band include the S(0), S(1) and S(2) pure rotational lines of H<sub>2</sub>, atomic fine structure lines of \[SI\], \[FeI\], \[SiII\], \[FeII\] and molecular lines of OH and H<sub>2</sub>O (GH04). The Formation and Evolution of Planetary Systems (FEPS) Spitzer Legacy Science Project (Meyer et al 2005) aims to study the evolution of gas in disks by targeting a carefully chosen sample of about 40 nearby solar-type stars for high resolution (R$``$ 700) spectroscopic observations by the IRS (Houck et al. 2004). The sample was selected mainly on the criteria that they span an age range of 3-100 Myr, and that they be nearby ($`<140`$ pc) so that the expected weak line fluxes may be detectable with the IRS. A secondary criterion was that they show some infrared excess emission indicative of the presence of dusty disks. To some extent, we favored sources with strong X ray luminosities, since the X rays heat the putative gas and intensify the resultant line luminosities. This will be the most comprehensive survey to date of gas in transition and post-accretion systems in order to characterize its dissipation and to place limits on the time available for giant planet formation. In this paper, we present high resolution IRS data and detailed modeling results for one of the first sources observed through the FEPS gas program, HD 105. HD 105 is a young ($`30`$ Myr) solar-type star 40 pc from the Sun with an infrared excess indicative of a dust disk (Meyer et al. 2004). The star has completed its active accretion phase and the dust continuum is characteristic of an optically thin, debris disk. The youth and proximity of the star, the presence of dust (which may imply associated gas), the relatively low dust continuum luminosity that allows the detection of weak gas lines, and the presence of X-ray flux from the central star (a source of gas heating) make HD 105 a promising candidate for the detection of mid-infrared gas lines. Spitzer spectroscopic observations of HD 105, however, did not detect any gas emission lines, but instead set stringent upper limits on the line fluxes. The line flux upper limits are used to set limits on the gas surface density and mass in HD 105. Assuming that the gas distribution has an inner hole of radius $`r_{i,gas}`$ greater than about 0.5 AU, we find that there is very little gas in the disk beyond the inner hole, suggesting the end of the main planet-building epoch for gas giants in these regions. Debris disks have been observed around early (e.g., 49 Cet, A1V) and late-type (e.g., AU Mic; MIV) stars and are often characterized by their fractional infrared luminosities, $`f=L_{IR}/L_{bol}`$, which range from $`10^510^3`$. HD 105 is thus a somewhat luminous debris disk, with a fractional infrared luminosity, $`f=L_{IR}/L_{bol}4\times 10^4`$. The disk may have evolved past its main planet formation stage, and is young enough to perhaps be a “transition object” with some residual, detectable gas. Though gas is readily detected in young disks (ages $``$ a few Myrs, and $`f0.01`$, e.g. Dent et al. 2005), most debris disks do not appear to have detectable amounts of gas. Notable exceptions are the disks around 49 Cet($`f10^3`$) and $`\beta `$ Pictoris($`f10^3`$). CO emission has been observed from the J$`=32`$ (Dent et al. 2005) and the J$`=21`$ (Zuckerman, Forveille & Kastner 1995) transitions from the disk around 49 Cet, and ISO detected H<sub>2</sub> $`28\mu `$m emission as well (Thi et al. 2001). In $`\beta `$ Pic gas emission from metal atoms and ions orbiting in a disk has been reported, most recently by Brandeker et al (2004). They and Thebault & Augereau (2005) derive gas masses in the disk of $`0.10.4`$ M. This amount of gas is insufficient to brake the ions against radiation pressure; both groups estimate $`40`$ M is needed for braking. Chen et al. (2004) obtain upper limits on warm gas in $`\beta `$ Pic of about 11 M, although these Spitzer observations of the pure rotational lines of H<sub>2</sub> are not sensitive to the presence of cold ($`<50`$ K) or atomic gas. The directly detected gas suggests gas depletion in debris disks relative to the dust, in the few instances where gas has been observed. HD 105 is the first debris disk observed in our FEPS H<sub>2</sub> program aimed at studying gas dispersal in disks as they evolve. As we shall show below, depending on the gas temperature, Spitzer is able to detect gas masses $`110`$ M for sources as close as HD 105. The outline of the paper is as follows. We describe the source (§2), and discuss the analysis of the Spitzer observations (§3). We then describe the gas and dust disk modeling (§4). Our results are discussed in §5 and the summary and conclusions presented in §6. ## 2 Source Description HD 105 is a G0 V spectral type star (Houk 1978) located at an Hipparcos distance of 40 $`\pm `$1 pc (Perryman et al 1997). Based on its observed space motion and corroborating age diagnostics, Mamajek et al. (2004) conclude that it is a likely member of the Tuc-Hor association. The membership suggests an age of about 30 Myr for HD 105. There are various other indicators of the young age of this dwarf: its Li I $`\lambda `$6707 equivalent width (e.g. Cutispoto et al. 2002) is similar to the $``$50 Myr-old members of the IC 2602 and 2391 clusters (Randich et al. 2001), and it is stronger than that of the 120 Myr-old Pleiades stars (Soderblom et al. 1990); HD 105 has an active chromosphere as revealed by the detection of Ca II H and K emission (Henry et al. 1996) and its X-ray luminosity ($`L_\mathrm{X}=2.4\times 10^{29}`$ ergs s<sup>-1</sup>) suggests youth (Metanomski et al. 1998, Wichmann et al. 2003). Based on all these age indications, we assigned an age of 30$`\pm `$10 Myr for HD 105 (Meyer et al. 2004). Early ISO observations showed that HD 105 displays infrared excess emission at 60 and 90 $`\mu `$m (Silverstone 2000). Our recent Spitzer measurements detect no excess emission at 24$`\mu `$m, confirm the excess at 70 $`\mu `$m, and detect dust emission even at 160 $`\mu `$m (Meyer et al. 2004, hereafter M04). B. Mazin (private communication) reports a 3$`\sigma `$ upper limit to the 350 $`\mu `$m continuum flux of 2.1 Jy. We have searched the GALEX database at the position of HD 105 and found a source offset from the HD 105 position by about 3.2” in the NUV, but only 0.09” in the FUV, whereas the typical positional uncertainty is about 1”. However, there is no other GALEX source within 30” of HD 105. Figure 1 shows the Kurucz model SED and the observed SED, including the infrared, submillimeter, and GALEX data. As expected for a young star, the source has FUV excess. Meyer et al. (2004) presented early models of the dust distribution implied by the far infrared excesses. The lack of detectable excess emission at wavelengths shorter than 35 $`\mu `$m implies very little circumstellar dust in the inner ($`15`$ AU) disk. We discuss these early models and other possible model fits in §4. ## 3 Observations and Analysis ### 3.1 Observations and Data Reduction We obtained 9.9 - 37.2$`\mu `$m spectra of HD 105 on December 14, 2003 UT using the Infrared Spectrograph (IRS, Houck et al. 2004) onboard the Spitzer Space Telescope (Werner et al. 2004). The spectra we present here were obtained with the two echelle modules ($`R=700`$) covering 9.9 - 19.5$`\mu `$m with the Short High (SH) detector array, and 18.9 - 37.2$`\mu `$m with the Long High (LH) array. We also acquired low resolution observations ($`R=80`$), which were used to account for background flux levels (discussed below). The spectra were acquired using the standard Staring Mode AOT, in which the telescope is nodded to place the star at two positions along each slit, resulting in four separate pointings (two for SH, and two for LH). For the SH observations, we used integration times of 31.5 seconds for each Data Collection Event (DCE), and cycled eight times for a total on-source exposure time of 504 seconds. For LH we used the 14.7 second integration time, also cycled eight times, for a total exposure time of 235 seconds. High accuracy blue peak up ($`13.318.7\mu `$m) was performed on the star to ensure a pointing accuracy within 0<sup>′′</sup>.4. The brightness profile of HD 105 in the peak up image agrees with that expected from the point spread function. Individual DCEs were processed through the SSC pipeline (version S10.5.0) resulting in Basic Calibrated Data (BCD) products to which basic detector calibrations, dark current subtraction, cosmic ray detection, integration ramp fitting, and reduction to individual, two-dimensional slope images in units of electrons/second/pixel have been applied. Dark current measurements taken with the LH module within 12 hours of the HD 105 observations were used to mitigate the effects of outlier pixels whose dark currents vary on timescales of one to several days, rather than using standard dark current subtraction which employs darks taken over a wider separation in time. The number of LH outlier pixels (high dark current pixels not representative of expected gaussian noise) is less than 6% of the 128$`\times `$128 pixels on the array. However, they may dominate the noise within each echelle order if untreated. Corrections were applied to pixels flagged as anomalous in the pipeline, due to cosmic-ray saturation early in the integration, or having been preflagged as unresponsive. For these pixels we applied cubic spline interpolation technique, relying on 1.5 resolution elements on either side of the dead pixels in the dispersion direction. This signal reconstruction method has been successfully verified with observations of spectral standards exhibiting resolved and unresolved emission lines, particularly for the LH array. Individual 1-dimensional spectra were extracted from two dimensional images flat–fielded with observations of the zodiacal background. We used the offline SSC software to extract these spectra using the full width of each echelle order. Spectral response and flux calibration corrections were applied to these 1-dimensional spectra. The response functions and flux calibration are based on similarly processed and spatially flatfielded observations of photometric standards HR 6688 (K2III), HR 6705 (K1.5III), HR 7310 (G9III), and HR 2194 (A0V), and the MARCS-code stellar atmosphere models tailored for these stars (Decin et al. 2004). The response correction of the short wavelength end of the SH data ($`9.912.0\mu `$m) of HD105 relied solely on HR 2194, in order to minimize any possible discrepancies in the strength of the SiO fundamental band in the K giants. The 16 individual spectra for each module were next median combined on a spectral order basis. Each order was trimmed, removing $``$5% of the data at the blue ends where the throughput response of the arrays is lowest. The signal-to-noise in the continuum is lower than the signal-to-noise that would be inferred from repeatability among the 16 spectra. This suggests that the signal-to-noise is limited by residual errors due to fringing, uncertainty in the spectra response function and anomalous pixels among others. The errors in absolute fluxes, which here include both the star and zodiacal background, are thought to be $``$ 20 %. Although we did not obtain off-source exposures with SH or LH for sky subtraction, approximate corrections were applied, based on the background estimator available in the SPOT planning tool. These fluxes are provided in MJy/sr, which we then converted to Jy using the solid angles subtended by each of SH and LH apertures (1.25 $`\times `$ $`10^9`$ sr and 5.82 $`\times `$ $`10^9`$ sr, respectively). Since the SPOT background estimate is derived from measurements made with low spatial resolution, we also compared the extracted sky spectrum obtained with the IRS near HD 105 at low spectral resolution. We found agreement to within 20% of the fluxes provided by the background estimator between 15 and 40 $`\mu `$m when scaled by the appropriate solid angles. Figure 2 shows the high resolution spectrum of HD 105. Figure 3 highlights the regions around the strongest emission lines expected from our models (Gorti & Hollenbach 2004). The S/N is approximately 12 in the 16 – 18$`\mu `$m range, where the noise is computed as the 1-$`\sigma `$ RMS spread in the ratio between the sky-subtracted IRS spectrum and the adopted Kurucz photospheric model over this essentially featureless region of the continuum (excluding the 16.9 – 17.1 $`\mu `$m range where we searched for evidence of the H<sub>2</sub> S(1) line). Similarly, we estimate a S/N ratio of approximately 10 in the 26.5 – 28.5um range. The S/N ratio becomes very low at $`\lambda >35\mu `$m, where the LH throughput is at its lowest. ### 3.2 Analysis To estimate upper limits of unresolved lines, we fit the spectrum using a Levenberg-Marquardt (LM) algorithm assuming a Gaussian for the line profile and a second order polynomial for the continuum. The central wavelength of the line was fixed in the fit, as was the width of the Gaussian to match the instrumental line profile. To fit the baseline locally, we used a wavelength range of $`\pm `$ 0.5$`\mu `$m centered on each feature (see Table 1). Five sigma upper limits to the line flux for each non–detection were derived taking the local RMS dispersion in the averaged continuum over two pixels per resolution element. We show examples of hypothetical 5$`\sigma `$ lines in Figure 3. Several anomalous pixels are evident in the spectra (e.g. at 12.45 and 16.56 $`\mu `$m) that were not flagged in the data reduction discussed in §3.1. Since these do not correspond to any expected emission lines, this suggests that residual systematic effects may be present in the spectra shown. Table 1 gives the upper limits to the line fluxes of H<sub>2</sub>, \[SI\], \[FeII\], and \[SiII\] lines to estimate upper limits of gas mass in this system. ## 4 Modeling the Gas and Dust in the Disk Around HD105 ### 4.1 Simple Estimate of the Upper Limit to Warm H<sub>2</sub> Mass Before presenting detailed models of gas/dust disks around HD 105, which can provide estimates of upper limits to the total gas mass in atomic and molecular form and at a (model–determined) range of temperatures, we calculate here a simple upper limit to the warm H<sub>2</sub> mass assumed to be at a constant temperature $`T`$ in the HD 105 disk. We use the upper limits on the H<sub>2</sub> S(0) 28$`\mu `$m and S(1) 17$`\mu `$m line fluxes to directly set limits on the mass of warm H<sub>2</sub> at temperature $`T`$ in HD 105. The H<sub>2</sub> S(0) and S(1) transitions have critical densities much lower than the gas densities in disks with gas masses $`3\times 10^5`$ M<sub>J</sub>, the extreme lower limit for H<sub>2</sub> line detection by Spitzer of HD 105 or other sources at 40 pc (see below). Therefore, for any detectable disk or even for disks with H<sub>2</sub> masses somewhat below detectability, the lower rotational levels (i.e., J=0-3) of H<sub>2</sub> are in LTE. With LTE and optically thin assumptions we calculate the mass $`M(H_2)`$ of molecular gas at temperature $`T`$ and at the distance (40 pc) of HD 105 that will produce a line flux $`F`$ for both the S(0) and S(1) transitions. $$M(H_2)2.8\times 10^5\left(\frac{F_{S(0)}}{10^{22}\mathrm{Watts}\mathrm{cm}^2}\right)\left(1+\frac{T}{85\mathrm{K}}\right)e^{510\mathrm{K}/T}M_J$$ (1) $$M(H_2)7.4\times 10^7\left(\frac{F_{S(1)}}{10^{22}\mathrm{Watts}\mathrm{cm}^2}\right)\left(1+\frac{T}{85\mathrm{K}}\right)e^{1020\mathrm{K}/T}M_J$$ (2) Utilizing these equations and the upper limits for the line fluxes given in Table 1, we find upper limits on the H<sub>2</sub> gas mass of 4.6 M<sub>J</sub> at 50 K, 3.8$`\times 10^2`$ M<sub>J</sub> at 100 K, and 3.0$`\times 10^3`$ M<sub>J</sub> at 200 K. The S(0) line flux sets the limit at 50 and 100 K, whereas the S(1) line flux sets the limit at 200 K. The S(2) line does not set useful limits at these low temperatures, where it is very weak. Note that assuming long Spitzer integrations that set H<sub>2</sub> S(1) flux limits of order $`10^{22}`$ W cm<sup>-2</sup>, the minimum H<sub>2</sub> mass detectable at 40 pc via the S(1) line is $`3\times 10^5`$ M<sub>J</sub>, achieved when the gas is $`T1020`$ K (see Eq. 2). Equations (1 and 2) simply relate the flux in an H<sub>2</sub> line to the mass of gas at a particular temperature. In the following section, we apply much more sophisticated models which actually calculate the temperature distribution in a disk, and relate the fluxes to the total mass of gas distributed in the disk. ### 4.2 Thermal/Chemical Modeling of Gas and Dust #### 4.2.1 Dust modeling The Spectral Energy Distribution (SED) obtained from existing data in the literature and new Spitzer observations (see Figure 1) has been modeled using the dust disk models of Wolf & Hillenbrand (2003). The initial dust disk model for HD105 was presented in M04 and we only give a brief description of these previous results here. The observed infrared continuum excess cannot be uniquely fit by any one particular dust model but by a range of dust parameters. M04 assumed for simplicity that the dust is composed of astronomical silicates with a surface density distribution $`\mathrm{\Sigma }(r)r^0`$. The model fits were relatively insensitive to the exponent in the radial density distribution and the outer disk radius $`r_{o,dust}`$, since much of the dust emission at wavelengths shorter than the peak emission arises from dust near the inner radius, $`r_{i,dust}`$. A grain size distribution of $`n(a)a^s`$ with $`s=3.5`$ was chosen, representative of regions with grain shattering, and providing somewhat better fits to the data than a single grain size. With such a distribution, most of the dust mass is in the largest particles, while the source of the infrared emission derives from grains with the minimum size $`a_{min}`$ (which holds most of the dust area). We define “dust” to be particles with sizes less than $`a_{max}=1`$ mm, and this definition sets the dust mass for a model fit. The three main parameters which determine the fit to the observed SED are then the inner radius $`r_{i,dust}`$, the minimum grain size $`a_{min}`$, and the dust mass $`M_{dust}`$. Roughly speaking, for a given $`r_{i,dust}`$ and $`a_{min}`$, the dust mass sets the total solid angle subtended by dust grains, and therefore sets the ratio $`L_{IR}/L_{bol}`$ of the dust IR luminosity to the stellar bolometric luminosity (an observed quantity which then determines the dust mass). M04 found acceptable fits with $`r_{i,dust}=32`$ AU and $`a_{min}=8\mu `$m, and with $`r_{i,dust}=45`$ AU and $`a_{min}=5\mu `$m. These best fit $`r_{i,dust}`$ and $`a_{min}`$ corresponded to dust masses of 9$`\times 10^8`$ and $`4\times 10^7`$ M. Since the results presented in M04, we have performed a more extensive parameter study of the best $`\chi ^2`$ fit, and have found that the minimum in $`\chi ^2`$ corresponds to even larger $`a_{min}=21`$ $`\mu `$m and lower $`r_{i,dust}=19`$ AU than explored in M04, and to somewhat lower dust mass $``$$`7.5\times 10^8`$ M. We found that quite good fits could be obtained for 13 AU $`<r_{i,dust}45`$ AU. The inner dust radius cannot be smaller than 13 AU because the dust becomes too hot, and produces excess emission at $`\lambda 35`$ $`\mu `$m. We do not emphasize this SED fitting procedure here because we have found, and show below, that for all acceptable dust models, there is too little dust surface area to affect the gas chemistry, heating, or cooling. Therefore, the gas spectrum is independent of the dust in the case of HD 105, and mainly depends on $`r_{i,gas}`$ and the gas surface density $`\mathrm{\Sigma }_0`$ at $`r_{i,gas}`$, as we discuss below. #### 4.2.2 Gas/dust disk models We apply the thermal/chemical disk models of Gorti & Hollenbach (2004) to study HD 105. In these models, a central star and the interstellar radiation field illuminate a gas and (optically thin) dust disk extending from $`r_i`$ to $`r_o`$. The gas and dust are heated by the radiation field (the gas is particularly sensitive to the UV and X-ray radiation), and the gas and dust temperatures are calculated in separate thermal balance equations. The vertical density structure and chemistry is self-consistently computed by imposing thermal balance, steady-state chemistry and pressure equilibrium. The assumption in the dusty regions is that the gas and small dust particles are well mixed, so that their density ratio does not vary vertically (no settling of small dust particles). However, as we shall show, there is so little dust in HD 105 that the gas emission lines do not depend on the dust vertical (or radial) distribution. We consider various gas heating sources such as gas-dust collisions, X-rays, stellar and interstellar FUV radiation, and exothermic chemical reactions. The cooling is mainly by molecular rotational and atomic and ionic fine structure emission. In summary, the inputs to the model include the stellar parameters, the interstellar field, $`r_i`$ and $`r_o`$, the surface density distribution of the gas, the dust size distribution, and the dust-to-gas mass ratio (which often is held fixed). Our main input variable generally is the gas surface density distribution. The output is the vertical density, chemistry, and temperature distribution as a function of $`r`$, and the resultant line intensities of various ionic, atomic and molecular species. In turn, these line intensities can be compared with line observations to constrain the physical parameters in the disk, such as the gas surface density distribution (or in this case provide upper limits of gas surface densities). We first consider a standard case, our best fit dust distribution discussed in §4.2.1, with $`r_{i,gas}=r_{i,dust}=19`$ AU, i.e., gas and dust co-exist spatially. However, we have found in this standard case, and in other SED-fitting cases we have tested with co-existing dust where $`r_{i,dust}=1345`$ AU, that the dust has no effect on the gas properties. There is too little surface area to appreciably affect the cooling (through gas-dust collisions), the heating (through the grain photoelectric mechanism), or the H<sub>2</sub> chemistry (through catalysis of H<sub>2</sub> on grain surfaces). Therefore, we have also considered a number of other models with different $`r_{i,gas}`$, but we have not included dust since its effect is minimal. As we shall show in the next section, the line fluxes from the gas originate from the inner regions, $`r_{i,gas}<r<r_{w,gas}`$, where $`r_{w,gas}`$ is defined such that 90% of the line luminosity is generated from $`r_{i,gas}`$ to $`r_{w,gas}`$. Considerable (gas) opacity to the stellar photons occurs at $`r_{i,gas}`$, so that $`r_{w,gas}`$ is typically only slightly ($`10`$%) larger than $`r_{i,gas}`$, leaving the shielded outer gas beyond $`r_{w,gas}`$ cold and unemissive. Therefore, there can be very large amounts of cold gas in these outer regions, depending on the gas outer radius. The small amount of emission from this cold gas is below Spitzer line flux constraints. The IRS on Spitzer, however, is very sensitive to the warm gas just between $`r_{i,gas}`$ and $`r_{w,gas}`$. Because the emitting region has such small radial extent, we assume the gas surface density $`\mathrm{\Sigma }_{gas}(r)=\mathrm{\Sigma }_0`$ is a constant, independent of $`r`$ in the emitting region. We then vary $`\mathrm{\Sigma }_0`$ and compute the line fluxes. The observed line flux limits are then used to constrain $`\mathrm{\Sigma }_0`$ and the gas mass $`M_w`$ between $`r_{i,gas}`$ and $`r_{w,gas}`$. If the gas extends far beyond $`r_{w,gas}`$, the total mass of the gas in the disk can be estimated from $$M_{gas}=\frac{2\pi \mathrm{\Sigma }_0r_{i,gas}^2}{2\alpha }\left[\left(\frac{r_{o,gas}}{r_{i,gas}}\right)^{2\alpha }1\right]\mathrm{for}\alpha <2,$$ (3) where the gas surface density is given by $`\mathrm{\Sigma }_{gas}(r)=\mathrm{\Sigma }_0(r_{i,gas}/r)^\alpha `$ for $`r_{i,gas}<r<r_{o,gas}`$. Generally, $`\alpha `$ is assumed to be $`01.5`$. Note that for $`\alpha <2`$ and $`r_{o,gas}>>r_{i,gas}`$, considerable cold (hidden) gas mass is located at the outer radius $`r_{o,gas}`$. The gas emission line fluxes then depend on only two main parameters, $`r_{i,gas}`$ and $`\mathrm{\Sigma }_0`$, in the case of HD 105 where dust does not affect the gas properties. The line fluxes vary with these parameters for the following reasons. X-rays and UV photons tend to penetrate and heat a fixed column of gas. Therefore, larger inner holes increase the mass of gas relative to the total disk mass that is affected by X-ray and UV photons. However, at the same time, the stellar radiation flux falls as $`1/r^2`$ and the flux incident on the inner edge of the disk decreases as $`r_{i,gas}`$ increases. The total energy intercepted by the disk does not vary appreciably with changing $`r_{i,gas}`$, because the scale height tends to scale as $`r`$ so that the gas disk subtends a fairly constant solid angle. However, the declining flux means a drop in the gas temperature (for fixed density), and the relative strengths of emission lines change. Similarly, $`\mathrm{\Sigma }_0`$ controls the gas density at $`r_{i,gas}`$, and increasing $`\mathrm{\Sigma }_0`$ changes the gas temperature and the relative strengths of the lines. We note that at small enough $`r_{i,gas}`$ the lines will become undetectable by Spitzer regardless of $`\mathrm{\Sigma }_0`$. This arises because the mass and surface area of the emitting gas decline with $`r_{i,gas}`$. Optically thin LTE line luminosities are proportional to mass, while optically thick lines are proportional to the emitting area. At very small $`r_{i,gas}`$ and when $`\mathrm{\Sigma }_0`$ is raised to extremely high values, the lines cannot exceed their (optically thick) blackbody limits, the temperature is finite, and the small surface area and mass drive the beam diluted fluxes below detectability. We consider in the next section a range of $`r_{i,gas}`$ from 0.5 to 100 AU. We have chosen cases of 13, 19 and 45 AU because they correspond to the minimum, the best fit, and an approximation of the maximum $`r_{i,dust}`$ allowable from the SED modeling. However, we also consider the possibility that the gas does not co-exist with the dust. In each case, we vary $`\mathrm{\Sigma }_0`$ and determine the variation in the line fluxes with $`\mathrm{\Sigma }_0`$ at the specified inner radius. ## 5 Results and Discussion ### 5.1 Gas Disks with Inner Holes We first discuss the results of the standard case where we have included the best fit dust population and $`r_{i,gas}`$ = $`r_{i,dust}`$ = 19 AU. Figure 4 shows that model line luminosities initially increase as the gas mass surface density increases at the inner radius. The flux from the \[SiII\] line is the strongest followed by the \[SI\], \[FeII\], and the H<sub>2</sub> S(1) line. The \[SiII\] and \[SI\] lines are seen to plateau at high surface densities as they reach their optically thick, blackbody limits and as Si<sup>+</sup> begins to recombine at high density. The dotted lines give the corresponding Spitzer upper limits presented in Table 1. The intersection of the (dotted line) upper limits with the (solid line) model curves for a given species marks the critical column density $`\mathrm{\Sigma }_{0,crit}`$ above which the line should have been detected. With this particular choice of $`r_{i,gas}`$, the line providing the strongest constraint is the \[SI\] line, which gives $`\mathrm{\Sigma }_{0,crit}0.14`$ gm cm<sup>-2</sup>. It is instructive to look at the details of a fiducial model (the standard case at the critical surface density for that case, $`\mathrm{\Sigma }_{0,crit}=0.14`$ gm cm<sup>-2</sup>) to understand the typical chemistry, temperature structure, and heating and cooling agents. The upper left panel of Figure 5 shows the gas temperature and the temperature of the smallest ($`a_{min}=21`$ $`\mu `$m) dust grains at the midplane as a function of $`r`$. The gas density at $`r_{i,gas}`$ in the midplane is about $`10^9`$ cm<sup>-3</sup>. We have suppressed the data from 19 AU to 19.04 AU (a region where the gas goes from predominantly atomic hydrogen to mostly molecular hydrogen and where where soft, $`0.5`$ keV X rays are absorbed), because of complicated behavior there which cannot be discerned at this graphic resolution. This small region does not significantly contribute to the line spectrum, since most of the emission arises from the much greater mass of gas that lies between 19.04 AU and 22 AU. At the inner disk edge (19 AU, not shown), the gas is atomic and fairly warm ($`500`$ K), but it very quickly cools to about 150 K at 19.04 AU because of the rapid rise in abundance of cooling molecular species such as CO. One sees that beyond 19.04 AU the temperature drops from $`150`$ K to about 70 K at 22 AU. The chemistry changes slowly in this region. The hydrogen is mostly H<sub>2</sub>, formed by the reaction of H with H<sup>-</sup> and on grain surfaces. We have found that eliminating grains does not change the H<sub>2</sub> abundance appreciably (it is nearly entirely molecular anyway). The atomic H abundance is quite high because of the inefficiency of the H<sup>-</sup> and grain processes. Most of the sulfur is atomic, most of the silicon and iron is singly ionized. The carbon is mostly in CO, and all remaining gas phase oxygen is in atomic O. Note that when $`r=`$ 19.1 AU, there is already a column $`N10^{22}`$ cm<sup>-2</sup> of hydrogen between the central star and that point. Because the densities are very high, and the UV field low, there is considerable H<sub>2</sub>, C, CO, S, Si, and Fe in the surface regions, and their column provides considerable opacity to the stellar UV photons via photoionization and photodissociation processes. The dust surface area per hydrogen atom is small, so that dust extinction is not important, and the opacity is entirely due to gaseous species. Similarly, the grain surface area is too small to effectively heat the gas by the grain photoelectric heating mechanism, or to heat or cool the gas by gas-grain collisions. The gas-grain cooling could not be shown in the lower right panel because it is many orders of magnitude below the graph. We have also run alternate SED-fitting models of allowable $`r_{i,dust}`$ and found that dust is not important in HD 105 for modeling the gas in the disk because of its low abundance and surface area. The main heating mechanisms for the gas in the emission zone are the heating by gaseous absorption of X-rays from the central star and of the stellar optical and UV photons which photoionize and photodissociate atoms, ions and molecules. The lower energy X-ray photons are absorbed nearest to the surface and, because of their higher cross sections for absorption, lead to the highest X-ray heating there. The more energetic X-rays penetrate further, and provide gas heating at greater depths. In the bulk of the emissive zone, the heating by optical and UV photons is dominated by the photoionization of S (which provides most of the electrons) and the photodetachment of electrons from H<sup>-</sup>. Note that this latter process requires only stellar optical photons and not the higher energy UV photons. However, this process requires H<sup>-</sup>; the H<sup>-</sup> is produced by electrons which arise from the photoionization of S and Si; and these photoionizations do require UV photons. The cooling is mainly by CO mid J rotational transitions ($`200400`$ $`\mu `$m), \[SI\] 25 $`\mu `$m, and \[OI\] 63 $`\mu `$m. Beyond 22 AU, the cooling is dominated by CO mid J transitions, as the gas cools below 70 K. If the gas disk extends much beyond 30 AU, these CO transitions become detectable if the CO does not freeze onto the cold grain surfaces. To date, no CO observations of HD 105 have been made. We note that the cooling by CO and \[OI\] occurs at longer wavelengths ($`\lambda >38`$ $`\mu `$m) than accessible by the IRS on Spitzer, but that these transitions may be detectable by instruments on the future Stratospheric Observatory for Infrared Astronomy (SOFIA) and by the Herschel Observatory. Direct heating of the gas by absorption of stellar photons heats the gas to $`T70`$K in the inner regions (19-22 AU) of the gas disk. The gas vertical scale height is proportional to $`T^{1/2}`$, and the warm gas intercepts about 20% of the stellar radiation due to its flared nature. In many wavelength bands the gas opacity is significant. Therefore, there is significant heating by this mechanism, but it tends to occur at the inner rim ($`r_{i,gas}`$), where the stellar fluxes are highest and least attenuated by the gas. The heating at the inner rim makes it expand vertically, enhancing its ability to intercept photons and to shield the outer disk. Dullemond, Dominik & Natta (2001) have discussed this effect for dusty disks where the opacity is provided by the dust particles. We find in the fiducial case that heated region extends radially about 3 AU beyond $`r_{i,gas}`$ to $`r_{w,gas}22`$ AU. Of order 90% of the luminosity of the infrared lines studied here is emitted in this narrow annulus. Because of this fact, the gas spectra from the model disks around HD 105, where dust plays no factor, depend only on $`r_{i,gas}`$ and the gas surface density $`\mathrm{\Sigma }_0`$ there, as discussed above. We now present the results of an identical search of $`\mathrm{\Sigma }_0`$ parameter space for a variety of $`r_{i.gas}`$ from 0.5 to 100 AU, assuming dust is unimportant. Table 2 lists the surface density limits $`\mathrm{\Sigma }_{0,crit}`$ as determined from each line. Associated with each of these lines and $`\mathrm{\Sigma }_{0,crit}`$ are particular values for $`r_w`$ and $`M_w`$, the mass of the warm gas between $`r_{i,gas}`$ and $`r_w`$. These are also listed in Table 2 for the most sensitive line. Note that the warm gas masses are somewhat less than the simple analytic expression given in §4.1 for H<sub>2</sub> lines produced in $`T100`$ K gas. Spitzer can detect masses somewhat smaller than $`10^2`$ M<sub>J</sub> in HD 105 (the limit derived from H<sub>2</sub> S(1) for $`T100`$ K gas) because other lines such as \[SI\] are predicted to be stronger, and, in the cases of small $`r_{i,gas}`$, the gas is warmer than 100 K. The total mass of gas (warm and cold) in any given model depends on the power law of the gas surface density distribution $`\alpha `$ and on the outermost extent $`r_{o,gas}`$ of the gas. Let us assume that the region of gas giant formation extends to about 40 AU, based on the solar system example. We take one of the more extreme values of the power law of the gas surface density distribution, $`\alpha =0`$, which maximizes the gas mass in the outer shielded zones. The upper limit (derived from the most sensitive line \[SI\]) to the total gas mass in this extreme case of low $`\alpha `$ is then about 1.9 M<sub>J</sub> for r$`{}_{i,gas}{}^{}=0.5`$ AU, 0.2 M<sub>J</sub> for r$`{}_{i,gas}{}^{}=1`$ AU, and 0.1 M<sub>J</sub> for r$`{}_{i,gas}{}^{}=520`$ AU, where we basically extrapolate the upper limit on the gas surface density measured at $`r_{i,gas}`$ to 40 AU. Therefore, for $`r_{i,gas}0.5`$ AU, there is insufficient gas in HD 105 at this time to feed the formation of gas giants. Table 2 shows that the upper limits to the gas surface density at $`r_{i,gas}`$ are not very sensitive to $`r_{i,gas}`$ for 1 AU $`r_{i,gas}40`$ AU. Larger $`r_{i,gas}`$ tends to produce more mass of gas heated by the stellar photons, but, because of the dilution of the stellar flux, also tends to lead to lower characteristic gas temperature. These effects counterbalance each other to some extent, resulting in similar line fluxes. However, for $`r_{i,gas}<1`$ AU, $`\mathrm{\Sigma }_{0,crit}`$ rises steeply. The heating mechanisms only penetrate to a relatively constant column of $`N_w10^{22}`$ cm<sup>-2</sup> so the mass of heated gas scales roughly as $`r_{i,gas}^2`$, assuming the vertical scale height scales with radius. Therefore, the mass (or area for optically thick lines) of heated gas goes down as we move inward, and so the gas temperature must rise appreciably in order for the gas to be detected. Since the gas temperature tends to rise with increasing density, due to the collisional de-excitation of the upper levels of cooling transitions, we require substantially more surface density to raise the temperature sufficiently to overcome the loss of warm gas mass and area. Figure 6 visually shows how the critical surface density $`\mathrm{\Sigma }_{0,crit}`$ depends on $`r_{i,gas}`$. We see that \[SI\] 25 $`\mu `$m is the most sensitive Spitzer line for constraining the gas surface density in gas disks with little dust, like HD 105. If $`r_{i,gas}`$ is sufficiently small ($`0.5`$ AU), the gas mass in the disk is essentially unconstrained by the Spitzer observations because a very small mass and surface area of inner gas shields the outer regions, leaving them too cold to be detected at this time. Clearly, if a line detection is made by Spitzer IRS, knowledge of the parameter $`r_{i,gas}`$ is extremely helpful if the line flux is to be converted into $`\mathrm{\Sigma }_{0,crit}`$ and $`M_w`$. Therefore, followup observations by higher spectral resolution instruments will be very useful, since resolved line widths can be translated into Keplerian velocities and, hence, $`r_{i,gas}`$. ### 5.2 Discussion Assuming that $`r_{i,gas}0.5`$ AU, the Spitzer upper limits to the gas lines fluxes in \[SI\] 25 $`\mu `$m, \[SiII\] 35 $`\mu `$m, \[FeII\] 26 $`\mu `$m, and H<sub>2</sub> S(1) 17 $`\mu `$m provide upper limits to the gas surface density at $`r_{i,gas}`$ which would indicate less than a Jupiter mass of gas is present in the gas giant planet-forming zone which extends to 40 AU. It is interesting that the most sensitive indicators of gas are not the H<sub>2</sub> lines, but fine structure lines (in particular \[SI\] 25 $`\mu `$m) of less abundant species (see Gorti & Hollenbach 2004). These upper limits indicate that in this region of possible gas giant planet formation, either a giant planet has already formed or it never will: the era of significant gas accretion onto protoplanets is over for this 30 Myr old system. The main caveat is that if the gas surface density distribution, unlike the dust, extends to the innermost regions ($`r_{i,gas}<0.5`$ AU) of the disk, then significant (i.e., $`>1`$ M<sub>J</sub>) amounts of cold gas could be hidden in the 5-40 AU region where gas giant planets may still be accreting, without violating our observational constraints on the line luminosities. However, if the gas indeed extends to $`<0.5`$ AU, then the gas is likely accreting onto the stellar surface. If $`M_d`$ is the mass of gas from the stellar surface to 40 AU, then the accretion rate onto the central star is given $`\dot{M}3\times 10^9(\alpha /0.01)(M_d/`$M<sub>J</sub>) M yr<sup>-1</sup>, where $`\alpha `$ is the standard alpha viscosity parameter. Such high accretion rates would produce diagnostics, such as large UV or near-IR excesses, H$`\alpha `$ emission lines, or other indicators of accretion-driven winds, which are not observed. Therefore, we regard this possibility as unlikely. The gas limits in the 10-40 AU region may also be relevant to theories of the formation of the outer giants such as Neptune and Uranus. A generic problem for these systems is that they heat dynamically (i.e., induce high random velocities) the much smaller objects which provide potential coalescent collisions that enable them to grow. The higher velocities reduce the gravitational focusing, and therefore reduce their growth rate (Levison and Stewart 2001). Without some kind of dynamical cooling mechanisms, it is difficult for them to form in situ within the lifetime of the solar system. Gas drag is one possibility, but it requires many Jupiter masses of gas. Our upper limits on the amount of gas indicate that, after 30 Myr in HD 105, there is not nearly enough gas drag to have an appreciable effect (see, e.g., Goldreich et al. 2004) at this time. Takeuchi and Artymowicz (2001), Klahr & Lin (2001), and Takeuchi & Lin (2002) showed that a small amount of gas (and the effect of stellar photons on the gas and dust) can sculpt the dust morphology and create an inner hole. Therefore, the sharp inner dust hole at $`19`$ AU in HD 105 need not reflect the presence of giant planets, but may be produced in a planet-less gas/dust disk around HD 105. These authors show that $`M_{gas}>110M_{dust}`$ produces such effects, and that when $`M_{gas}>100M_{dust}`$ the dust grains cannot migrate effectively relative to the gas. Using our upper limits for the case where the gas and dust are co-spatial, we find that we can only constrain the gas mass to dust mass ratio in HD 105 to values of $`<1000`$, and that therefore we cannot effectively constrain the effects of gas on the dust dynamics. In other words, the sharp inner rim of dust implied by the IR continuum SED of HD 105 could be produced by a small amount of gas below our upper limits. An alternate hypothesis is that in the absence of gas, a giant planet is preventing dust from the outer debris disk from reaching the inner part of the system, as discussed in M04. Kominami & Ida (2002, 2004) and Agnor & Ward (2002) discuss the effects of a small amount of gas in the terrestrial zone (1-5 AU) on the resulting formation of terrestrial planets. If $`M_{gas}>>10^2`$ M<sub>J</sub> in the terrestrial zone for tens of millions of years, then lunar and Mars-sized planetary embryos feel the dynamical friction of the gas, circularize their orbits, and never collide to form Earth-mass planets. On the other hand, if $`M_{gas}<<10^2`$ $`M_J`$, then the embryos are on eccentric orbits, and collide to form Earth-sized or larger orbits, but with eccentricities substantially larger than the Earth. The suggestion is that the Earth may have formed with roughly 10<sup>-2</sup> M<sub>J</sub> of gas in the terrestrial zone for tens of millions of years. The terrestrial zone ($`0.33`$ AU) in HD 105 is inside $`r_{i,dust}`$ and is currently quite dust-free. Our models show that if the gas extends all the way in to $`r_{i,gas}<0.5`$ AU, then we are not sensitive to gas mass in the terrestrial zone, and therefore cannot set useful limits on this process. If for some reason the gas had an inner radius of 0.5 – 1 AU, we can set limits of about $`2\times 10^2`$ to $`2\times 10^3`$ M<sub>J</sub> for the mass of gas from this inner radius to 5 AU. In this case, the limits are close to the critical value of $`10^2`$ M<sub>J</sub>, and indicate perhaps insufficient gas to prevent large, Earth-sized planets from forming. Spitzer may be able to set even more stringent limits on this process in other sources. If sufficient small dust is mixed with the gas in the terrestrial zone, then the gas in the models is hotter, and smaller gas masses can be detected. Finally, are these upper limits surprising in the context of theoretical calculations that have been performed on the likely dispersal mechanisms of the gas? The current kinematic evidence suggests that HD 105 formed in a small group of tens of stars (see §2), which likely lacked O or early B-type stars. Assuming that HD 105 was not exposed to high UV fluxes from nearby massive stars, the main dispersal mechanism for the outer disk would likely be photoevaporation of the gas by the central star, and for the inner disk viscous accretion and spreading of the gas would dominate (Hollenbach, Yorke & Johnstone 2000, Clarke et al. 2001, Matsuyama et al. 2003). Unfortunately, it is difficult to answer this question because of the lack of self consistent photoevaporation models which treat not just the EUV (i.e., $`h\nu >13.6`$ eV) photons (see Hollenbach et al. 1994), but also the less energetic stellar photons as well as the X rays from the young central star. Gorti & Hollenbach (in preparation) are developing such models, and their preliminary results show that the less energetic photons rapidly (in less than 10 Myr) disperse the gas outside of about 30-50 AU. The EUV photons can potentially remove gas outside of about 1-5 AU, but the evolution of the EUV luminosity and the radiative transfer of these photons as they try to penetrate the protostellar winds in the early stages of star formation are not well determined. A recent paper by Alexander, Clarke & Pringle (2005) suggests rather high escaping EUV luminosities which may rapidly ($`<10`$ Myr) remove the outer gas. Once the outer disk is removed, viscosity removes the inner gas on timescales which are roughly 0.1 Myr (0.01/$`\alpha _v)(r/`$10 AU), where $`\alpha _v`$ is the turbulent viscosity parameter in the standard “$`\alpha `$” disks. The value of $`\alpha _v`$ is typically 10<sup>-2</sup> if the Balbus & Hawley (1991) magneto-rotational instability, or MRI, is operant. The MRI instability requires a minimal level of ionization, which all our relatively low mass models meet. Assuming that MRI is active, the viscous timescales inside the photoevaporation region (which lies at $`>330`$ AU) are of order 0.03 – 0.3 Myr –extremely short! On the other hand, it is not totally certain that this instability would be fully active (Chiang, Fischer, & Thommes 2002). Therefore, we conclude that these upper limits are not surprising, but they do set constraints on the rather poorly known dispersal mechanisms. ## 6 Summary and Conclusions One of the goals of the Formation and Evolution of Planetary Systems (FEPS) Spitzer Legacy project is to measure the evolution and dispersal of gas in the planet-forming regions ($`0.540`$ AU) of disks around solar-type stars of ages 3–100 Myr. We report here our first carefully reduced and analyzed high spectral resolution data taken by the IRS instrument on the Spitzer Space Telescope. This paper illustrates our method of modeling the data to obtain constraints on the gas surface density distribution and mass. We eventually plan to obtain data on about 40 nearby stars in this age range to look for variation with age and other stellar properties. The data presented here are for the source HD 105, a $`30`$ Myr old G0 star at a distance of 40 pc with a known IR excess arising from a circumstellar dust disk orbiting at $`r_{i,dust}13`$ AU. The derived upper limits to the H<sub>2</sub> S(0) 28 $`\mu `$m, H<sub>2</sub> S(1) 17 $`\mu `$m, H<sub>2</sub> S(2) 12 $`\mu `$m, \[SI\] 25 $`\mu `$m, \[FeII\] 26 $`\mu `$m, and \[SiII\] 35 $`\mu `$m lines are given in Table 1. The H<sub>2</sub> upper limits directly place limits on the mass of warm gas in the disk: $`M(\mathrm{H}_2)4.6`$ M<sub>J</sub> at 50 K, $`3.8\times 10^2`$ M<sub>J</sub> at 100 K, and 3.0$`\times 10^3`$ M<sub>J</sub> at 200 K. This can be compared with the roughly 10<sup>-3</sup> M<sub>J</sub> of gas detected around $`\beta `$ Pictoris, a 10-20 Myr old A5 star at a distance of about 19 pc (Brandeker et al 2004). It can also be compared with the recent UV absorption measurements and analysis of AU Mic, an M1 star in the $`\beta `$ Pictoris Moving Group at a distance of 9.9 pc and with a similar age as $`\beta `$ Pic (Roberge et al 2005). They find an upper limit to the H<sub>2</sub> mass of about $`2\times 10^4`$ M<sub>J</sub>. It appears that even in the inner ($`<30`$ AU) regions not well probed by CO observations, the gas is largely dissipated in these three sources which span the age range 10-30 Myr, and which span the stellar types from M1 to A5. Detailed thermal/chemical models of HD 105 were constructed and compared with the Spitzer observations to obtain further constraints on the gas mass and surface density. These models calculate the gas temperature, chemistry, and vertical structure self-consistently and predict line fluxes which depend largely on the gas inner radius, $`r_{i,gas}`$, and the gas surface density $`\mathrm{\Sigma }_0`$ there. We show that most of the gas emission arises in a thin inner rim, extending from $`r_{i,gas}`$ to $`r_{w,gas}`$, heated by stellar optical, ultraviolet, and X-ray photons. The upper limits on the \[SI\] 25 $`\mu `$m and \[SiII\] 35 $`\mu `$m fine structure lines provide the strongest constraints on the gas surface density $`\mathrm{\Sigma }_0`$ at the inner rim. If the gas inner radius is comparable to the observationally constrained dust inner radius, $`r_{i,dust}13`$ AU, we show that the Spitzer upper limits on the line fluxes limit $`\mathrm{\Sigma }_0`$ to $`0.2`$ gm cm<sup>-2</sup> for $`r_{i,gas}`$ between 10 and 40 AU. In this case, the total mass of gas (cold and warm) in the planet forming region between 10 and 40 AU is constrained to be less than 0.1 M<sub>J</sub> (assuming a constant surface density between $`r_{i,gas}`$ and 40 AU). The gas may not co-exist with the dust, however. We show that even if the putative gas extends inward to $`r_{i,gas}=0.5`$ AU, the Spitzer upper limits set tight constraints on the gas surface density and mass. The upper limits on the gas surface density are $`\mathrm{\Sigma }_{0,crit}`$ = 3.43, 0.36, and 0.12 gm cm<sup>-2</sup> for $`r_{i,gas}`$ = 0.5, 1, and 5 AU, respectively. The total mass of gas in the gas giant planet-forming region out to 40 AU depends on how we extrapolate the gas surface density from the inner radius. If we assume a gas surface density power law $`\mathrm{\Sigma }r^{3/2}`$, which is often assumed in disks, then the total gas mass in these three cases is limited to about $`10^2`$, 3$`\times 10^3`$, and $`10^2`$ M<sub>J</sub>, respectively. An extreme assumption, allowing for the most hidden cold gas mass, would be a constant surface density with radius. In this case, we obtain total (warm and cold) gas mass limits of 1.9, 0.2, and 0.07 M<sub>J</sub>, respectively. In summary, assuming that $`r_{i,gas}>0.5`$ AU and any reasonable gas surface density distribution, there is less than a Jupiter mass of total gas in the gas giant planet forming region out to 40 AU. Given likely temperature distributions produced in our models, Spitzer is unlikely to be able to detect total gas masses less than about 10<sup>-2</sup> M<sub>J</sub> for disks around relatively nearby ($`30pc`$) low mass stars. If the gas extends to $`r_{i,gas}0.5`$ AU, the Spitzer upper limits set no useful constraints on the gas mass, because gas lines become undetectable even with extremely large gas masses. In this case, the small inner rim which absorbs the heating photons from the star has too little mass (for optically thin lines) and too little surface area (for optically thick lines) to provide detectable Spitzer emission, regardless of the magnitude of $`\mathrm{\Sigma }_0`$. The outer gas is effectively shielded from the heating photons, and significant cold gas mass could exist in the planet forming regions. We argue that this case is unlikely, however, since the gas would likely extend all the way to the stellar surface, and viscous accretion onto the central star would lead to observational diagnostics such as near infrared or UV excess, H$`\alpha `$ emission, and signs of winds generated by the accretion process (which are not observed). We therefore conclude that the Spitzer upper limits imply low upper limits to the gas surface density and mass in the 0.5-40 AU region around HD 105. We also note that, like HD 105, many debris disks will have too little dust to affect the gas spectra; therefore, a similar analysis of Spitzer upper limits on line fluxes will result in similar conclusions. As discussed in §5, the limits are not sufficiently stringent to constrain the potential effects of gas on dust dynamics. They set interesting limits on the mass of gas which might affect terrestrial planet formation only in the ad hoc case where the inner gas radius is of order 0.5 – 1 AU. However, the limits do set interesting constraints for giant planet formation. The upper limits to the gas mass obtained here are too small to enhance the buildup of gas poor outer giants such as Uranus or Neptune, and too small to allow for gas giants to form from the gas reservoir after this time ($`30`$ Myr). We acknowledge support from NASA’s Spitzer Space Telescope Legacy program, which has supported our group, the Formation and Evolution of Planetary Systems Legacy team. We thank the rest of the FEPS team for their efforts in making the FEPS project successful, and for their help in obtaining, analyzing, and interpreting the data discussed in this paper.
warning/0506/astro-ph0506384.html
ar5iv
text
# A near-IR spectrum of the DO white dwarf RE J0503-285Based on observations collected at the European Southern Observatory, Chile. ESO No. 072.D-0362 ## 1 Introduction In accord with a theoretical prediction made several decades ago (Schatzman schatzman58 (1958)), the atmospheres of white dwarfs are observed to be dominated either by hydrogen (DAs) or by helium (DO/DBs). Unimpeded, their high surface gravities lead to the settling out of heavier elements on timescales of mere days to months (e.g. Dupuis et al. dupuis93 (1993)). In the standard theory of single star evolution the distinction between the two compositions is made at birth. If a star leaves the AGB during a period of quiescent hydrogen shell burning it evolves onto the hydrogen dominated cooling channel. Alternatively, if a thermal pulse occurs as the star evolves off the AGB, much of the remaining hydrogen is incinerated and nuclear processed material from deeper layers (e.g. C and O) is dredged up to the surface leading to a hydrogen deficient object (e.g. Iben et al. 1983, Herwig et al. herwig99 (1999)). A number of empirical results are broadly supportive of this hypothesis. For example, the observed number ratios of hydrogen rich to hydrogen-deficient central stars of planetary nebulae (2:1; Mendez mendez91 (1991)) and DAs to DOs at T$`{}_{\mathrm{eff}}{}^{}>40000`$K (7:1; Fleming et al. fleming86 (1986)) are roughly consistent with theoretical expectations (Iben & Tutukov iben84 (1984)). The observed transformation of the hydrogen deficient PG1159 central star of the planetary nebula, Lo4, into a Wolf Rayet object and back over a period of months (Werner et al. werner92 (1992)) and similarities in the measured abundances of the elements C and O in the atmospheres of Wolf Rayet (C/He$`0.2`$ and O/He$`0.05`$, Koesterke & Hamann koesterke97 (1997)) and PG1159 stars (C/He$`0.10.6`$ and O/He$`0.0050.1`$, Dreizler & Heber dreizler98 (1998)) confirm an evolutionary link between these two classes of object. Furthermore, the abundance patterns observed in the atmospheres of the DO white dwarfs ($`120000>`$T$`{}_{\mathrm{eff}}{}^{}>45000`$K, C/He$`0.0010.01`$, O/He$`<0.001`$; Dreizler dreizler99 (1998)) suggest they are the descendents of the PG1159 stars ($`180000>`$T$`{}_{\mathrm{eff}}{}^{}>65000`$K), gravitational settling having reduced the abundances in the former. Despite detecting $``$110 hot DA white dwarfs (T$`{}_{\mathrm{eff}}{}^{}>25000`$K), the extreme-ultraviolet surveys of the ROSAT WFC and EUVE unearthed only one new DO white dwarf, RE J0503-285. An analysis of IUE and Voyager data and the optical identification spectrum revealed an effective temperature and surface gravity of T$`{}_{\mathrm{eff}}{}^{}70000`$K and log g =7.5-8.0 respectively (Barstow et al. barstow94a (1994)). These values are consistent with more recent determinations utilising more refined synthetic spectra and/or higher S/N optical data (T$`{}_{\mathrm{eff}}{}^{}=72660_{6289}^{+2953}`$K, log g $`=7.50_{0.15}^{+0.13}`$, Barstow et al. barstow00 (2000); T$`{}_{\mathrm{eff}}{}^{}=70000`$K, log g=7.50, Dreizler & Werner dreizler96 (1996)). Somewhat counter-intuitively to it’s detection as an EUV source, of the 14 DOs studied in detail by Dreizler & Werner (dreizler96 (1996)) and Dreizler (dreizler99 (1998)), RE J0503-318 is one of the most metal rich, with C/He$`0.005`$, O/He$`0.0005`$, N/He$`10^5`$ and Si/He$`10^5`$. Nickel has also been detected in the photosphere at an abundance of Ni/H$`10^5`$ (Barstow et al. barstow00 (2000)). However, no convincing evidence of the presence of iron is found in any of the spectral datasets obtained to date, setting an upper limit on its abundance of Fe/H$`<10^6`$. To an extent this is an intriguing result when one notes that the cosmic abundance ratio of Fe:Ni is 18:1, that Fe is observed to be more abundant than Ni in the atmospheres of hot DA white dwarfs by factors between 1 to 20 and that self-consistent model calculations taking into account gravitational settling and radiative levitation predict an apparent Fe:Ni ratio of $``$1:3 (e.g. Dreizler dreizler99 (1998)). Barstow & Sion (barstow94b (1994)) report evidence of episodic massloss in a series of IUE spectra of RE J0503-285. It is plausible that the observed pattern of abundances may be reflective of this process. Alternatively, the abundance pattern may be related to the hypothetical dredging up of nuclear processed material from deeper layers caused by the late-thermal pulse, as has been proposed to explain the Fe deficiency observed in a number of PG1159 stars (e.g. Miksa et al. miksa02 (2002)). However, as the mass of RE J0503-285 maybe somewhat lower than the canonical DO value of $`0.59\pm 0.08`$M (Dreizler & Werner dreizler96 (1996)) it is a possibility that binary evolution (e.g. Vennes et al. vennes98 (1998)) has influenced or continues to interfere with the photospheric composition. Indeed, as discussed by Bragaglia et al. (brag90 (1990)) it is possible that close binary evolution produces preferentially non-DA white dwarfs with hydrogen-deficient envelopes. The lack of any prior published detailed examination of this particularly interesting object at near-IR wavelengths has motivated us to obtain, during a recent study of DA white dwarfs, a JHK spectrum of RE J0503-285 to investigate external factors possibly influencing its evolution and photospheric composition e.g. the presence of a cool low mass companion. We present here the results of our analysis of this data. ## 2 Observations ### 2.1 Data acquisition A low resolution near-IR spectrum of RE J0503-285 was obtained using the ESO New Technology Telescope (NTT) and the Son-of-Isaac (SOFI) infrared instrument on 2003/12/10. The sky conditions at the La Silla site were good on this night with seeing typically in the range 0.6”-1.0”. We used the low resolution spectroscopic mode ($`\lambda /\delta \lambda 950`$ with the 0.6” slit), in which coverage of the wavelength ranges $`0.951.64\mu `$m and $`1.532.52\mu `$m is provided by the “blue” and the “red” grism respectively. The observations were undertaken using the standard technique of nodding our point source target back and forth along the spectrograph slit in an ABBA pattern. The total integration times used for the blue and the red grism setup were 1080s and 2880s respectively. To facilitate the removal of telluric features from the target spectra and to provide an approximate flux calibration, a standard star (HIP28999) was observed immediately after the science integrations. This was carefully chosen to lie within $`0.1`$ airmasses of the white dwarf. ### 2.2 Data reduction We have used software routines in the STARLINK packages KAPPA and FIGARO to apply standard reduction techniques to our data. Further details of these procedures are given in Dobbie et al. (dobbie05 (2005)). Here we re-iterate that any features intrinsic to the energy distribution of the standard star were identified by reference to a near-IR spectral atlas of fundamental MK standards (Wallace et al. wallace00 (2000), Meyer et al. meyer98 (1998), Wallace & Hinkle wallace97 (1997)) and were removed by linearly interpolating over them. Furthermore, the flux levels of the data were scaled to (1) achieve the best possible agreement between the blue and the red spectrum of the white dwarf in the overlap region between $`1.531.64\mu `$m and (2) obtain the best possible agreement between the spectral data and the J, H and K<sub>S</sub> photometric fluxes for RE J0503-285 derived from the 2MASS All Sky Data Release Point Source Catalogue magnitudes (Skrutskie et al. skrutskie97 (1997)) where zero magnitude fluxes were taken from Zombeck (zombeck90 (1990)). The reduced spectrum and 2MASS fluxes are shown in Figure 1. ## 3 Analysis of the data ### 3.1 Model DO white dwarf spectra We have generated a He+C+H synthetic white dwarf spectrum with abundances C/He$`=0.005`$ and H/He$`=10^5`$, at the effective temperature and surface gravity shown in Table 2. We have used the latest versions of the plane-parallel, hydrostatic, non-local thermodynamic equilibrium (non-LTE) atmosphere and spectral synthesis codes TLUSTY (v200; Hubeny hubeny88 (1988), Hubeny & Lanz hubeny95 (1995)) and SYNSPEC (v48; Hubeny, I. and Lanz, T., private communication). The calculation included a full treatment of line blanketing and used state-of-the-art model atoms. In brief, the HeII ion incorporated the 19 lowest energy levels, where the dissolution of the higher lying levels was treated by means of the occupation probability formalism of Hummer & Mihalas (hummer88 (1988)), generalised to the non-LTE situation by Hubeny, Hummer & Lanz (hubeny94a (1994)). Lines originating from the four lowest levels were treated by means of an approximate Stark profile (Hubeny et al. hubeny94b (1994)) during the calculation of the model structure and in the spectral synthesis step. The CIV ion incorporated levels up to n=14 where levels $`9\mathrm{n}14`$ were each represented by a superlevel. Where available, photoionization cross sections were obtained from TOPbase. Transitions between the lowest level and those with n$`5`$ were represented by Stark profiles. The synthetic energy distribution has been normalised to the V magnitude of RE J0503-285 and convolved with a Gaussian to match the resolution of the SOFI spectrum. This is shown overplotted on the observed data in Figure 1. It is worth noting here that our modelling indicates that at the effective temperature of this white dwarf, the colours V-K, J-H and H-K are rather weak functions of T<sub>eff</sub>. ### 3.2 The search for a cool companion We have examined Figure 1 for significant differences in the overall shape or level between the observed and synthetic fluxes which can be consistent with the presence of a cool companion. Further, we have searched for specific features in the spectrum typical of the energy distributions of M or L dwarfs e.g. K I and Na I absorption at 1.25$`\mu `$m and 2.20$`\mu `$m respectively, CO at 2.3$`\mu `$m and H<sub>2</sub>O centred on 1.15, 1.4 and 1.9$`\mu `$m. However, no convincing evidence for such has been found. Hence, we have added empirical models for low mass stellar and substellar objects, full details of which are given in Dobbie et al. (dobbie05 (2005)), to the white dwarf synthetic spectrum and compared these composites to the IR data to set an approximate limit on the spectral type of a putative cool companion. The fluxes of the empirical models have been scaled to a level appropriate to a location at d=10pc using the 2MASS J magnitude of each late-type object and the polynomial fits of Dahn et al. (dahn02 (2002)) and Tinney et al. (tinney03 (2003)) to the M<sub>J</sub> versus spectral type for M6-M9 and L0-L8 field dwarfs/brown dwarfs respectively. These fluxes have been further reduced by a factor 1.4, corresponding to the rms dispersion in the M<sub>J</sub> versus spectral type relationship of Tinney et al. (tinney03 (2003)). Subsequently, the fluxes have been re-calibrated to be consistent with the distance of the DO as derived from the measured V magnitude, the effective temperature and theoretical M<sub>V</sub> and radius from evolutionary models of pure-C core white dwarfs which include only a He layer mass of $`10^4`$M (Wood wood95 (1995)). As the effective temperature and surface gravity and hence our distance determination of RE J0503-285 are considerably less well constrained than those of the DA white dwarfs we have previously investigated (e.g. Dobbie et al. dobbie05 (2005)), for a conservative limit we have adopted the largest distance estimate consistent with Barstow et al.’s analysis and the error limits given therein (see Table 2). In fact the neglect of line blanketing arising from metals other than carbon present in the atmosphere of REJ0503-285 impacts on the temperature estimate in a such a way as to reinforce our cautious approach (e.g. see Barstow et al. barstow00 (2000)). Starting with L8 we have progressively added earlier spectral types to the synthetic white dwarf spectrum, until it could be concluded with reasonable certainty that the presence of a companion of that effective temperature or greater would have been obvious from our data, given the S/N. ## 4 Results ### 4.1 A limit on the mass of a pututive companion to RE J0503-285 The effective temperatures of very-low-mass stars and substellar objects remain a function of both mass and age on the timescale typical of the lifetime of the white dwarfs likely progenitor ($``$Gyrs). Although we can estimate the cooling age of RE J0503-285 using theoretical evolutionary models, as we don’t know with any certainty the mass and hence lifetime of its progenitor star (the initial-mass final-mass relationship is poorly constrained for white dwarfs of this mass and in any case may not apply to DOs) a robust estimate of the age of a putative associate of the DO is not possible. Therefore, to use our limit on spectral-type to constrain mass, we instead assume a range of ages broadly encompassing the likely value. We assign an approximate effective temperature to the spectral type limit shown in Table 3 using the polynomial fit described in Table 4 of Golimowski et al. (golimowski04 (2004)). Subsequently, we refer to the low mass stellar/substellar evolutionary models for solar metallicity of Baraffe et al. (baraffe03 (2003)), using cubic splines to interpolate between their points, to estimate corresponding mass at ages 1Gyr, 5Gyrs and 10Gyrs as shown in Table 3. We note that there are a number of examples of post common envelope white dwarf + main sequence binaries, where a low mass companion has survived a phase of common envelope evolution e.g. RE J0720-318, REJ1016-05, and RE J2013+400 and now interacts with the white dwarf to influence the composition of its photosphere. Our result argues strongly against the presence of an unresolved cool companion to this white dwarf with M$`>0.085`$M and we can now rule out that the pattern of abundances observed in the atmosphere of RE J0503-285 is influenced by ongoing interaction with such an object. However, our result does not exclude the possibility that the evolution of RE J0503-285 was affected by the interaction of the white dwarf’s progenitor star and a close companion (M$`>0.085`$M) during a common envelope phase from which the latter failed to emerge. Further, there remains the possibility of a close companion with a mass below our detection threshold. This may seem unlikely given that an extensive near-IR study of a sample of 371 white dwarfs, summarised by Farihi, Becklin & Zuckerman (2005), has revealed only two secondaries with spectral types later than M, GD165B (L4; Becklin & Zuckerman 1998) and GD1400B (L6; Farihi & Christopher 2004, Dobbie et al. 2005), the former at a separation a$``$120AU from its white dwarf primary. Nevertheless, there are persistent although perhaps controversial claims for L dwarf secondaries in some cataclysmic variables such as EF Eri (e.g. Howell & Ciardi 2001, Harrison et al. 2003). ### 4.2 Possible future work We point out that RE J0503-285 and the other DO white dwarfs are potential targets for Spitzer imaging to search for a mid-IR excess due to companions cooler than spectral type M. Our non-LTE model of REJ 0503-285 indicates the white dwarf flux at 10$`\mu `$m to be $``$0.04mJy. From the Spitzer spectrum of DENIS-P J0255-4700 (Roellig et al. roellig04 (2004)), which we determine to reside at d=5.3pc using the 2MASS magnitude (J=13.25$`\pm `$0.03) and the M<sub>J</sub> versus spectral type relationship of Tinney et al. (tinney03 (2003)), allowing for a 25% uncertainty in the absolute calibration of the mid-IR data and the uncertainties in late-type dwarf fluxes and the white dwarf distance discussed in Section 3.2, we estimate the flux level of an L8 companion to RE J0503-285 to be $``$0.004mJy. As the absolute accuracy of Spitzer photometry is expected to be $``$5%, it should be possible to extend this search to the regime of the coolest L dwarfs, corresponding to substellar masses. Further, improved line broadening theories specifically tailored to helium plasmas would offer the possibility of reducing the uncertainties on our estimates of the effective temperatures and surface gravities of DO white dwarfs and allow more robust determinations of their distances. ###### Acknowledgements. PDD, MRB, ANL and RN are supported by PPARC. This publication makes use of data products from the Two Micron All Sky Survey, which is a joint project of the University of Massachusetts and the Infrared Processing and Analysis Center/California Institute of Technology, funded by the National Aeronautics and Space Administration and the National Science Foundation. We thank the anonymous referee for useful comments which have improved this work.
warning/0506/cond-mat0506235.html
ar5iv
text
# Universal aspects of non-equilibrium currents in a quantum dot ## I Introduction and discussion The description of an out-of-equilibrium strongly correlated system is a long standing problem. Even in the simplest case where the system is in a steady state and its properties no longer change with time, the usual formalism of quantum statistical mechanics is inadequate. Theoretical understanding of such systems became all the more pressing with the recent spectacular progress in nanotechnology, which has made it possible to study the Kondo impurity, one of the best understood strongly correlated systems, in out-of-equilibrium conditions. The Kondo impurity was realized experimentally as a quantum dot, a tiny island of electron liquid, attached via two tunnel junctions to leads (baths or reservoirs of electrons) held at different electric (or chemical) potentials. This set-up allows an electric current to flow across the dot, and measurements of the current were carried out as a function of the potential difference $`V`$, the temperature $`T`$ and the magnetic field $`B`$ exp . When the dot carries a net spin in the Coulomb blockade regime, it can be modeled by a Kondo Hamiltonian with two channels $`\alpha =1,2`$, corresponding to the two leads, to which the spin of the dot, $`\stackrel{}{S}`$, couples qdot . The resonant tunneling through the dot (elastic co-tunneling) allows the electrons from each bath to jump on the dot and back to the same bath, leading to the formation of Kondo resonance around the Fermi level $`\mu _\alpha `$ in each lead. Further, electrons from one bath can jump on the dot and onto the other bath, giving “off-diagonal coupling” of the two channels to each other. With the matrix of couplings $`J_{\alpha ,\alpha ^{}}`$ and at zero magnetic field, the Hamiltonian is $`H=`$ $`{\displaystyle \underset{\alpha }{}}{\displaystyle \underset{\stackrel{}{k},a}{}}(ϵ_\stackrel{}{k}\mu _\alpha )c_{\alpha ,\stackrel{}{k},a}^{}c_{\alpha ,\stackrel{}{k},a}+{\displaystyle \underset{\alpha ,\alpha ^{}}{}}{\displaystyle \underset{\stackrel{}{k},\stackrel{}{k}^{},a,a^{}}{}}J_{\alpha ,\alpha ^{}}c_{\alpha ,\stackrel{}{k},a}^{}\stackrel{}{\sigma }_{a,a^{}}c_{\alpha ^{},\stackrel{}{k}^{},a^{}}\stackrel{}{S}.`$ (1) Here $`a`$ denotes the spin index $`a=\pm 1/2`$, and $`\stackrel{}{S}`$ is in the spin-1/2 representation. The process corresponding to off-diagonal coupling induces a current when the baths are held at nonzero potential difference, $`V=\mu _2\mu _1`$. The development of the Kondo resonance as temperature is lowered enables the system to overcome the Coulomb blockade, producing a significant increase of the conductance. The unitarity limit is reached as $`T0,V0`$. As we are interested in the universal properties of the system, we shall consider the model in the range $`T,VD_\alpha `$, where $`D_\alpha =D`$ are the bandwidths of the leads, each lead being considered a very large conductor (the bandwidths can be assumed to be the same for both channels). We are allowed therefore to carry out the standard steps (linearizing around the Fermi level, keeping only the $`s`$-wave component in the expansion of $`c_{\alpha ,\stackrel{}{k},a}`$ in spherical modes), to obtain a representation of each lead as a free electron gas on the half line consisting of left and right movers $`\psi _{\alpha ,L}(x),\psi _{\alpha ,R}(x),x0`$, interacting with the impurity localized at $`x=0`$. It will be convenient for us to “unfold” the baths, making left and right movers on the half line into right movers on the full line (defining $`\psi _{\alpha ,R}(x)=\psi _{\alpha ,L}(x),x0`$). See Figure 1. The field-theoretic Hamiltonian is then: $`H`$ $`=`$ $`i{\displaystyle \underset{\alpha }{}}{\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑x\psi _\alpha ^{}(x)\psi _\alpha (x)+{\displaystyle \frac{V}{2}}{\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑x(\psi _2^{}\psi _2\psi _1^{}\psi _1)+{\displaystyle \underset{\alpha ,\alpha ^{}=1,2}{}}J_{\alpha ,\alpha ^{}}\psi _\alpha ^{}(0)\stackrel{}{\sigma }\psi _\alpha ^{}(0)\stackrel{}{S}`$ (2) $`=`$ $`H_0+VH_z+H_I`$ where we work in units such that $`v_F=1`$. We also denote, $$H_z=\frac{1}{2}_{\mathrm{}}^{\mathrm{}}𝑑x(\psi _2^{}\psi _2\psi _1^{}\psi _1)$$ and $$H_I=\underset{\alpha ,\alpha ^{}=1,2}{}J_{\alpha ,\alpha ^{}}\psi _\alpha ^{}(0)\stackrel{}{\sigma }\psi _\alpha ^{}(0)\stackrel{}{S}.$$ The coupling of the baths to the quantum dot is parametrized by the Hermitian matrix $$J_{\alpha ,\alpha ^{}}=\pi \left(\begin{array}{cc}\lambda _d& \lambda \\ \lambda & \lambda _d\end{array}\right)_{\alpha ,\alpha ^{}}$$ (3) where $`\lambda _d`$ and $`\lambda `$ are real (the factor $`\pi `$ is introduced for later convenience). It is possible to diagonalize the matrix of coupling $`J_{\alpha ,\alpha ^{}}`$ by a change of basis in the channel space. In the simple situation of a single level quantum dot, described by the Anderson model, coupled to identical leads one naturally finds the relation $`\lambda _d=\lambda `$. This makes the matrix of coupling constants degenerate, one of the eigenvalues being zero. After diagonalizing it, the terms representing the interaction with $`\stackrel{}{S}`$ become that of a decoupled free fermion and a one-channel Kondo model. More generally, for $`\lambda _d\lambda `$, diagonalizing the matrix of couplings gives the interaction term of the usual 2-channel Kondo model. Out of equilibrium, for $`V0`$, diagonalizing the matrix of couplings does not lead to a simplification of the problem, since the out-of-equilibrium term $`VH_z`$ induces extra coupling between the new fermion fields: it is not invariant under change of basis. We will not perform this diagonalization here, in order to keep the term $`VH_z`$ simple. Also, we will consider the general case $`\lambda _d\lambda `$. In this paper, we are interested in studying the electric current as function of voltage and temperature in the steady state of this model. As in the usual Schwinger-Keldysh formulation of a non-equilibrium steady state Schwinger61 ; Keldysh65 , we will imagine coupling the dot to the leads at some time $`t_0`$ in the past when the system is in a thermal equilibrium state, then allowing the system to evolve till $`t=0`$ when the current is evaluated. One expects that after a transitory regime, as $`t_0\mathrm{}`$, the system will relax into a steady state with a constant current flowing through the dot. Many questions arise concerning this formulation. Probably the most obvious one is: Is the model sufficient to describe the establishment of a steady state current? Or does one have to contemplate additional relaxation mechanisms (certainly present in actual experiments) to absorb the continuous flow of energy of the electrons moving from the higher Fermi-level lead to the lower one? In the framework of real-time perturbation theory, a related (but not equivalent) question that one can answer is whether or not the “infrared” limit $`t_0\mathrm{}`$ exists for the integrals representing the perturbative coefficients in an expansion in $`\lambda `$ and $`\lambda _d`$. We develop the real-time Keldysh perturbation theory (in some ways similar to BazhanovLZ99 and KaminskiNG00 ), and use it to establish the convergence of every term of the perturbative series as the switch-on time $`t_0`$ is sent to minus infinity. This result is highly non-trivial. Real-time perturbation theory often gives divergences as the switch-on time is sent to minus infinity, unless a good relaxation phenomenon is included in the model. In equilibrium, this issue can be easily overcome: an infrared-divergent real-time perturbation theory for a system in equilibrium only means that the particular model we are considering does not have the proper relaxation mechanism. But with an additional interaction, however small, representing a proper relaxation mechanism, the system will have infrared convergent real-time perturbation theory. Then, for equilibrium models, this can be equivalently described by the always-infrared-convergent “imaginary time” perturbation theory (where the integrals in imaginary time are on a finite interval) coming from the description of the model using its equilibrium density matrix. There, the additional interaction representing relaxation can be sent to zero from the beginning. Out of equilibrium, however, there is no a priori steady-state density matrix description of steady quantities. Thus, if the real-time perturbation theory is infrared divergent, there is no simple way to describe steady state physics. This is very natural: in contrast to the equilibrium case, the relaxation mechanism is used not only to reach the steady state, but also to form it, since we need a continuous absorption of energy. Hence, such infrared divergences are far more pathological. If the limit of large negative switch-on time exists order by order, this superficially seems a good indication that the steady state is reached and that the model indeed describes the steady state (although, strictly speaking, one would still have to analyse the possible non-perturbative contributions if the perturbative series is asymptotic). But if no such limit exists, certainly more questions arise: does the model reach a steady state non-perturbatively (that is, divergences are an artifact of perturbation theory), or are other interactions necessary? These questions of course appeared in the literature before (see for instance WingreenS ; rg6 ). An interesting example is the case studied in rg6 , where it was shown that when the model we are discussing is put into a magnetic field, the limit $`t_0\mathrm{}`$ and $`\lambda ,\lambda _d0`$ do not commute, leading to divergencies in the perturbation theory as $`t_0\mathrm{}`$. Assuming that a steady state exists in the model, the correct result as $`t_0\mathrm{}`$ should then be non-perturbative, and it was partially evaluated under this assumption (the “zeroth order” was evaluated), without external thermal reservoir coupled to the dot. But the convergence of the perturbative series or the technical “way around” its divergencies as just described above do not guarantee that the results are describing the correct physical steady state. Indeed, if the model is believed to have a steady state, perturbatively or not, two questions should still be answered: Is there an element in the model playing the role of a good thermal reservoir to sustain the correct steady state in the model? And if not, is there a guarantee that a coupling to an external reservoir would not have an important effect? These last two questions have more bearing than it may seem. In the usual Schwinger-Keldysh formulation, one does not assume any exchange with a thermal bath while the system is evolved: one starts with a thermal equilibrium state, then turns on the coupling to the dot and lets the system evolve without thermal bath. This is certainly not the true physical situation; in fact, understanding how a thermal bath affects the evolution of a quantum system was amongst the main points of the study of Caldeira and Leggett CaldeiraL81 . We carry out the real-time perturbation theory both in equilibrium (that is, in the case $`V=0`$; note that the real-time formalism includes a transitory non-equilibrium region where the system relaxes to equilibrium) and out of equilibrium ($`V0`$), and show that it is the same phenomenon that makes the perturbative series infrared convergent in equilibrium and out of equilibrium. This phenomenon is a factorization at large time separation of the correlation functions involved in the perturbative coefficients; interpreting the integrals over time defining these coefficients in terms of physical processes, this signals a decoherence in time induced by the leads and suggests that the leads are good thermal baths. The convergent expression in equilibrium is indeed the right equilibrium density matrix, confirming that the leads themselves play the role of thermal baths for the dot degrees of freedom. The convergent expression out of equilibrium should then be “the right” steady state. Note that the factorization signaling decoherence in time occurs because the Hamiltonian for the leads is conformally invariant: indeed, in conformal field theory, a large time separation is a large distance separation, which, by locality, gives rise to factorization. Physically, this occurs because the separation between the energy levels of the baths is much smaller than all other scales in the problem and essentially energy-independent (although, as will become clear from our investigation, these two conditions may not be sufficient). Note that these ideas are not entirely new: in CallanT90 the results of Caldeira and Leggett for constructing a thermal bath coupled to a quantum mechanical system were re-interpreted as coupling a bulk conformal field theory in a disk to degrees of freedom on the boundary. Our results generalize this to the behavior of the impurity in the Kondo model. We are currently investigating how this can be further generalized (for example, what the general properties of the conformal field theory should be). Our proof also allows us to describe the steady-state physics in terms of a “steady state density matrix,” as conjectured by Hershfield Hershfield93 . The essential difference between the usual density matrix and the steady-state density matrix can be seen as a non-locality in the latter which captures the build-up of the steady state. Note that a proof of convergence to all order was developed in BazhanovLZ99 for a non-equilibrium free boson model with boundary interactions, but the arguments there were very model-dependent (and do not apply to the non-equilibrium Kondo model) and quite different from ours. In particular, our arguments have a much deeper physical meaning and scope, as explained above. Let us stress here that it is quite important to know that no external bath is required for reaching a non-equilibrium steady state in an impurity model. Indeed, this means that we can use the real-time formalism without addition of a coupling to an external bath in order to 1) study the perturbative scaling properties of the model as we did in this paper for the Kondo model (see below), 2) construct more or less explicitly, without strong assumption, the steady state as an eigenstate of the Hamiltonian, for instance using the the formalism of the “steady-state density matrix”; this eventually can give access to the infrared behavior of the model and to the integrability properties of the steady state (work in progress). The real-time formalism might not be the easiest way of trying to obtain this understanding, but it is probably the clearest, as it is the most closely related to the actual experimental situation. Other questions that need to be addressed in the study of this out-of-equilibrium steady state are: How will the Kondo effect, the quenching of the impurity spin as the temperature is lowered below the Kondo scale $`T_K`$, evolve in the presence of a current? Will new scales make their appearance? Which quantities are universal? To what extent the powerful ideas of the Renormalization Group (RG) apply there? Many interesting attempts were carried out, mainly perturbatively, to understand the flow of couplings as the cut-off (band width) $`D`$ is reduced rg1 ; rg2 ; rg3 ; rg4 ; rg5 ; rg6 ; rg7 . We shall pursue a different track and study a question related to the universality features of the model, namely: does a limit $`D\mathrm{}`$ exist? In this limit all results are universal. We shall establish that such a limit exists by running the RG equations “backwards”, referring to them in the field theoretic context, in the usual way, as the Callan-Symanzik equations. We shall deduce an out-of-equilibrium $`\beta `$-function carrying out the calculation directly in the steady state and will show that it is the same as the equilibrium $`\beta `$-function. This may not be too surprising since the singularity structure of the system usually does not depend on the state in which they are evaluated, so that the ground state and the highly excited steady state produce the same singularities. The finite parts of course are different. We show that only one scale arises, the Kondo temperature, $`T_K`$, and the current can be written in a universal form as a function of the ratios $`T/T_K,V/T_K,C`$, with $`C`$ an additional dimensionless parameter characterizing the asymmetry between $`\lambda `$ and $`\lambda _d`$ (in other words, specifying the RG trajectory). We carry out the computation of the current to two-loop order and verify these statements explicitly. We then use the RG arguments to re-sum the leading logarithms. Our results are valid in the regime where both the bias voltage and the temperature are smaller than the band width, and where the bias voltage or the temperature is larger than the Kondo scale: $`V,TD`$, and $`T_KV`$ or $`T_KT`$. In particular, we verify that there are no divergencies at $`T/V0`$ in the perturbative results. This means that to two-loop order, the voltage plays the role of a good infrared cutoff. From the RG analysis, we give a universal prescription, valid in this regime, for comparing the effect of the electric current to the effect of the temperature on the destruction of the Kondo cloud. We also examined the effect of a local magnetic field on the dot but, as expected, were unable to show that the perturbation series converges in this case. We will come back to a discussion of this case, in relation with the results of rg6 , in the last section of this paper. ## II Formulation of the problem and general considerations The Schwinger-Keldysh formulation Schwinger61 ; Keldysh65 . First formulation. We shall be interested in the electric current that passes from lead $`2`$ to lead $`1`$ across the quantum dot under the action of the potential diffence $`V`$. It can be calculated by evaluating the average of the current operator $`𝒥`$ with respect to a density matrix that has evolved over sufficiently long time from the initial non-interacting density matrix $$\rho _0=e^{\beta H_0}$$ under the action of the full evolution operator $$S^{(V)}(t_1,t_2)=e^{i(t_1t_2)H}=e^{i(t_1t_2)(H_0+VH_z+H_I)}.$$ (4) The operational meaning of this formulation is the following. The non-interacting leads are initially, say at time $`t_0`$, brought to thermal and chemical equilibrium at zero potential difference exchanging energy and particles with a common external reservoir at fixed temperature and chemical potential. The energy levels of lead 1 and lead 2 are filled up to the same energy (with thermal and particle fluctuations). Just after time $`t_0`$, they are separated from the external reservoir, then a potential difference $`V`$ is applied and the interaction is turned on. The application of the potential $`V`$ just after time $`t_0`$, as usual, causes a raising of the energy levels of lead 2 with respect to those of lead 1. For the clarity of the discussion below, it is worth being more precise here. One should imagine both leads having a continuum of available states from the bottom of their bandwidths with increasing energies (the energies grow in a continuous way for infinite leads, of course, so one should think about densities of states). At time $`t_0`$, the available states of the leads are filled up to equal energies. Then, just after time $`t_0`$, when the potential is applied, one shifts the energies of all states of lead 2 by, say, $`V/2`$ (towards higher energies), and the energies of those of lead 1 by $`V/2`$, without changing the occupations of the states. Hence, the levels of lead 2 are now filled up to a higher energy than those of lead 1. Since the reservoir is disconnected and the interaction is turned on, there is a current. The steady state current is obtained after an infinite time, which we will take to be time 0 (that is, we will take $`t_0\mathrm{}`$). In the equation (4), the raising of the energy levels and the turning on of the interaction strength seem instantaneous and simultaneous. But one can multiply both terms $`VH_z`$ and $`H_I`$ by a factor that smoothly increases from 0 at time $`t_0=\mathrm{}`$ to 1 at time $`0`$, for instance the factor $`e^{\mu t}`$, in order to implement a simultaneous adiabatic increase of both the potential and the interaction strength. Sending $`\mu 0`$ (the adiabatic increase occurring far in the past) gives the steady state. This is really what is understood in this formulation. It is not obvious, a priori, that this formulation represents adequately the usual experimental situation, where the leads and the quantum dot are always connected to a common thermal reservoir (but not a reservoir of electrons), even while the steady state is being reached. However, it is natural to think that the leads can themselves play the role of thermal reservoirs. As discussed in the introduction, this is indeed the case, and will be made more precise below. The average of an operator $`𝒪`$ in the steady state is then given by $$𝒪_{ss}=\underset{t_0\mathrm{}}{lim}\frac{\mathrm{Tr}\left[S^{(V)}(0,t_0)e^{\beta H_0}S^{(V)}(t_0,0)𝒪\right]}{\mathrm{Tr}\left[e^{\beta H_0}\right]}.$$ (5) The operators act in the Hilbert space for $`H_0`$ (which is a tensor product of the two-channel free massless fermion Hilbert space and of the impurity space) obtained by imposing asymptotically vanishing conditions for the fermion fields (correlation functions of fermion fields vanish at infinite distance from each other and from the dot). To be more accurate, we could start by taking the fermion fields on a line segment of length $`L`$ containing the dot, with some free boundary conditions; for instance, $`\psi _\alpha (L)=\psi _\alpha (L)`$ (this corresponds to the usual free boundary conditions when fermions are folded back on the half line), then send $`L`$ to infinity. The steady state would be obtained in the region $$L^1|t_0|^1V,T$$ (6) where the energy scale of switch-on, $`|t_0|^1`$, suffices to smear out the energy level spacing $`L^1`$. Second formulation. Another formulation can be given. The initial idea of this second formulation is that the current can be created not only by a shift of the energies of the states of leads 1 and 2 (coming from the application of an electric potential), but also by putting additional electrons in lead 2 and taking away electrons from lead 1. In order to implement this, one starts again, at time $`t_0`$, with the uninteracting leads, both connected to a common thermal and particle reservoir, and in themal and chemical equilibrium; but now the chemical equilibrium is not at potential difference 0, but rather at a potential difference $`V`$. The initial density matrix is then $$\stackrel{~}{\rho }_0=e^{\beta (H_0VH_z)}.$$ This potential difference shifts towards lower energies the states of lead 2 with respect to those of lead 1 by an amount $`V`$. But since there is equilibrium, the states of lead 1 and lead 2 are still filled up to the same energy. Note that then, as compared to the first formulation at time $`t_0`$, there are more available sates of lead 2 and less of lead 1 that are filled. Just after time $`t_0`$, the reservoirs are disconnected, then the potential is set to 0 and the interaction is turned on. The density matrix $`\stackrel{~}{\rho }_0`$ then evolves with the evolution operator at zero bias voltage $`\overline{S}(t_1,t_2)`$, $$\overline{S}(t_1,t_2)=S^{(V=0)}(t_1,t_2)=e^{i(t_1t_2)(H_0+H_I)}.$$ (7) Putting the potential to 0 has the effect of raising the energy levels of lead 2 with respect to those of lead 1 by an amount $`V`$, the same effect that occurs in the first formulation just after time $`t_0`$ when the potential is applied. In contrast, though, this brings us to a situation where the available states of leads 1 and 2 have exactly the same energies as in the first formulation at time $`t_0`$ (that is, at potential 0), but with more states filled in lead 2 and less in lead 1, so that the leads are filled up to unequal energies. This indeed implements having put additional electrons in lead 2 and extracted electrons from lead 1. With the interaction on and the reservoir disconnected, a current is created. Again, after an infinite time, the steady state should be reached. The current is then given by $$𝒥_{ss}=\underset{t_0\mathrm{}}{lim}\frac{\mathrm{Tr}\left[\overline{S}(0,t_0)e^{\beta (H_0VH_z)}\overline{S}(t_0,0)𝒥\right]}{\mathrm{Tr}\left[e^{\beta (H_0VH_z)}\right]}$$ (8) (for more general operators $`𝒪`$, see (32)). If the size of the bandwidth can be sent to infinity (when evaluating quantum averages of operators that give finite results in this limit), then the operational description above for the second formulation is equivalent to that of the first formulation, since then only the Fermi energies of leads 1 and 2 matter. In particular, raising the energy levels or filling states with electrons are exactly the same operation in this case. However, there is another difference between both formulations. In the second formulation, we can now think about putting a factor $`e^{\mu t}`$ for adiabatically increasing the interaction strength, but there is no such possibility for adiabatically increasing the potential (one would have to add the term $`(1e^{mut})VH_z`$ in the evolution Hamiltonian). In other words, for practical calculations, this second formulation naturally implies that the energy levels of lead 2 are raised (or the states are filled), with respect to those of lead 1, by an amount $`V`$ instantaneously, and that the interaction is then turned on adiabatically. This is to be contrasted with the first formulation, where both the potential difference and the interaction strength were understood as being simultaneously increased adiabatically. We will show below that both formulations are equivalent. Symmetry currents. The Hamiltonian $`H_0`$ is conformally invariant and has a large algebra of symmetries associated with it. It is a WZW “current algebra” of the symmetry currents (not to be confused with the physical current $`𝒥`$) and it will be convenient to carry out many of the calculations in terms of symmetry currents. Introduce the following operators, $`J_z`$ $`=`$ $`\pi :(\psi _2^{}\psi _2\psi _1^{}\psi _1):`$ $`\stackrel{}{J}_x`$ $`=`$ $`i\pi (\psi _2^{}\stackrel{}{\sigma }\psi _1\psi _1^{}\stackrel{}{\sigma }\psi _2)`$ $`\stackrel{}{J}_y`$ $`=`$ $`\pi (\psi _2^{}\stackrel{}{\sigma }\psi _1+\psi _1^{}\stackrel{}{\sigma }\psi _2)`$ $`\stackrel{}{J}_d`$ $`=`$ $`\pi :(\psi _2^{}\stackrel{}{\sigma }\psi _2+\psi _1^{}\stackrel{}{\sigma }\psi _1):.`$ (9) They form the following subalgebra of the $`su(4)_1`$ current algebra: $`[J_d^i(x),J_d^j(y)]`$ $`=`$ $`2i\pi \left(ϵ_{ijk}J_d^k(x)\delta (xy)\delta _{ij}\delta ^{}(xy)\right)`$ $`[J_x^i(x),J_x^j(y)]`$ $`=`$ $`2i\pi \left(ϵ_{ijk}J_d^k(x)\delta (xy)\delta _{ij}\delta ^{}(xy)\right)`$ $`[J_y^i(x),J_y^j(y)]`$ $`=`$ $`2i\pi \left(ϵ_{ijk}J_d^k(x)\delta (xy)\delta _{ij}\delta ^{}(xy)\right)`$ $`[J_z(x),J_z(y)]`$ $`=`$ $`2i\pi \delta ^{}(xy)`$ $`[J_d^i(x),J_x^j(y)]`$ $`=`$ $`2i\pi ϵ_{ijk}J_x^k(x)\delta (xy)`$ (10) $`[J_d^i(x),J_y^j(y)]`$ $`=`$ $`2i\pi ϵ_{ijk}J_y^k(x)\delta (xy)`$ $`[J_d^i(x),J_z(y)]`$ $`=`$ $`0`$ $`[J_x^i(x),J_y^j(y)]`$ $`=`$ $`2i\pi \delta _{ij}J_z(x)\delta (xy)`$ $`[J_y^i(x),J_z(y)]`$ $`=`$ $`2i\pi J_x^i(x)\delta (xy)`$ $`[J_z(x),J_x^i(y)]`$ $`=`$ $`2i\pi J_y^i(x)\delta (xy).`$ In terms of these currents the full Hamiltonian $`H=H_0+VH_z+H_1`$ can be expressed as follows: $`H_0`$ $`=`$ $`{\displaystyle \frac{1}{4\pi }}{\displaystyle _{\mathrm{}}^{\mathrm{}}}dx(:\stackrel{}{J}_d(x)\stackrel{}{J}_d(x):+:\stackrel{}{J}_x(x)\stackrel{}{J}_x(x):+:\stackrel{}{J}_y(x)\stackrel{}{J}_y(x):+:J_z(x)J_z(x):),`$ $`H_z`$ $`=`$ $`{\displaystyle \frac{1}{2\pi }}{\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑xJ_z(x)`$ $`H_I`$ $`=`$ $`\lambda _d\stackrel{}{J}_d(0)\stackrel{}{S}+\lambda \stackrel{}{J}_y(0)\stackrel{}{S}.`$ (11) The operator $`H_z`$ is the total (normalized) isospin $`z`$-component. <sup>1</sup><sup>1</sup>1One could also introduce the operator $`J_c=\pi (\psi _2^{}\psi _2+\psi _2^{}\psi _1)`$. Since the total (sum over the two channels) charge is locally conserved, the current $`J_c`$ commutes with all current operators introduced above, hence it decouples from the theory and can be identically set to 0. The electric current. We now turn to discuss in more detail the current across the quantum dot (we set the electric charge $`e=1`$), and express it also in terms of the symmetry currents. The electric current $`𝒥`$ is given by the operator measuring the difference between the fermion density on, say, the second channel just before hitting the impurity and the fermion density on the same channel just after hitting it: $$𝒥=\underset{ϵ0^+}{lim}\left(\psi _2^{}\psi _2(x=ϵ)\psi _2^{}\psi _2(x=ϵ)\right)=\frac{1}{2\pi }\underset{ϵ0^+}{lim}(J_z(ϵ)J_z(ϵ)).$$ (12) (we use here the “unfolded set-up”). Equivalently one can express the electric current as the rate of decrease of the charge on lead-2 (or increase on lead-1), $`𝒥`$ $`=`$ $`{\displaystyle \frac{d}{dt}}N_2={\displaystyle \frac{d}{dt}}N_1=i[H,H_z]`$ (13) $`=`$ $`\lambda \stackrel{}{J}_x(0)\stackrel{}{S}.`$ It is easy to see that the two definitions coincide. We rewrite the first definition using “impurity conditions” - operator relation inherited from boundary conditions. Boundary conditions, in general, are part of the equations of motion and lead to operator relations valid on the full Hilbert space of a boundary quantum field theory. They are often derived from the action of the model in the same way as one derives the equations of motion. From our view point, after “unfolding” the Kondo model, we have a model with an impurity instead of a boundary. As with boundaries, impurities give rise to “impurity conditions” which are part of the equations of motion and are operator relations valid on the full Hilbert space. In operator language (which is more convenient for our purposes), the impurity condition associated to a local operator $`𝒪(x)`$ can be written $$\underset{ϵ0^+}{lim}\left(_{\mathrm{}}^ϵ𝑑x+_ϵ^{\mathrm{}}𝑑x\right)[H,𝒪(x)]=_{\mathrm{}}^{\mathrm{}}[H,𝒪(x)].$$ Consider the impurity condition associated to the operator $`J_z(x)`$ with the model with Hamiltonian $`H`$ (2), $$\underset{ϵ0^+}{lim}\left(_{\mathrm{}}^ϵ𝑑x+_ϵ^{\mathrm{}}𝑑x\right)[H,J_z(x)]=_{\mathrm{}}^{\mathrm{}}[H,J_z(x)].$$ (14) On the left-hand side, only the free part $`H_0`$ of the Hamiltonian is involved, because the operator $`J_z(x)`$ is never at the site $`x=0`$ (and $`H_z`$ commutes with $`J_z(x)`$). Using the fact that with the free Hamiltonian $`H_0`$, $`J_z(x)`$ is a right-moving operator $`[H_0,J_z(x)]=i\frac{d}{dx}J_z(x)`$, and using the asymptotic conditions $`J_z(\mathrm{})=J_z(\mathrm{})`$, we have $$\underset{ϵ0^+}{lim}\left(_{\mathrm{}}^ϵ𝑑x+_ϵ^{\mathrm{}}𝑑x\right)[H,J_z(x)]=\underset{ϵ0^+}{lim}i(J_z(ϵ)J_z(ϵ)).$$ On the other hand, on the right-hand side of (14), since the integration is on the full interval, the free part of the Hamiltonian does not contribute. Only the impurity term, at $`x=0`$, contributes, and it gives $$_{\mathrm{}}^{\mathrm{}}[H,J_z(x)]=2i\pi \lambda \stackrel{}{J}_x(0)\stackrel{}{S}.$$ as expected. Having now discussed the system and the various operators describing it we turn to discuss in more detail the nature of non-equilibrium in the system. ## III Equilibrium vs. non-equilibrium Our model, a quantum impurity coupled to leads at different chemical potentials, describes a non-equilibrium situation – a current is flowing from one lead to another. What is the the precise meaning of this statement? In this section we show in what sense an out-of-equilibrium model differs from an equilibrium model. We begin by showing how the Keldysh formulation leads, when the system is in equilibrium, to the usual equilibrium density matrix description. We prove, in other words, the following: $$\frac{\overline{S}(0,\mathrm{})e^{\beta H_0}\overline{S}(\mathrm{},0)}{\mathrm{Tr}\left[e^{\beta H_0}\right]}=\frac{e^{\beta H_0}𝒫\mathrm{exp}\left(i_{i\beta }^0𝑑tH_I^{(0)}(t)\right)}{\mathrm{Tr}\left[e^{\beta H_0}𝒫\mathrm{exp}\left(i_{i\beta }^0𝑑tH_I^{(0)}(t)\right)\right]}=\frac{e^{\beta H|_{V=0}}}{\mathrm{Tr}\left[e^{\beta H_{V=0}}\right]}$$ (15) as an equation to hold when evaluated inside traces with insertion of any number of local operators at fixed positions. A local operator is, by definition, an operator depending on the position $`x`$ (in the sense that its commutator with the momentum operator is a derivative with respect to $`x`$), such that its commutator with the hamiltonian density at position $`y`$ is zero for $`xy`$. Note that local charges, for instance conserved charges of the Hamiltonian, are integrals of local operators, and are not local operators themselves. Hence, the limit (15) does not hold with insertion of local charges. This makes physical sense, since conserved charges are not expected to relax to their equilibrium values. Technically, one must remember that the density matrix is an operator with infinitely many matrix elements, hence any limit applied to it cannot be expected to converge to an object having the same properties (or to converge at all) independently from which subset of matrix elements we are looking at. This derivation is important for what follows, so we present it in some detail. Recall that $`\overline{S}(t_1,t_2)`$, Eq. (7), is the evolution operator at zero voltage. We now establish some useful identities. In the interaction picture with respect to $`H_0`$ we have, $`\overline{S}(0,t_0)e^{iH_0t_0}e^{\beta H_0}`$ $`=`$ $`𝒫\mathrm{exp}\left(i{\displaystyle _0^{t_0}}𝑑tH_I^{(0)}(t)\right)e^{\beta H_0}=e^{\beta H_0}𝒫\mathrm{exp}\left(i{\displaystyle _{i\beta }^{t_0i\beta }}𝑑tH_I^{(0)}(t)\right)`$ (16) where in the interaction picture $$H_I^{(0)}(t)=e^{iH_0t}H_Ie^{iH_0t}=\lambda _d\stackrel{}{J}_d(t)\stackrel{}{S}+\lambda \stackrel{}{J}_y(t)\stackrel{}{S}.$$ In (15) and in the last two expressions of (16), the symbol $`𝒫`$ indicates path-ordering in time: the operators are positioned from left to right with their time argument going from the lower integral limit to the upper integral limit. In the first occurrence in (16), integrals are ordered from $`0`$ on the left to $`t_0`$ on the right. In the second, the integration contour is from $`i\beta `$ on the left to $`t_0i\beta `$ on the right. On the other hand, we have, $$e^{iH_0t_0}\overline{S}(t_0,0)=𝒫\mathrm{exp}\left(i_{t_0}^0𝑑tH_I^{(0)}(t)\right).$$ (17) The Keldysh evolution is then (multiplying (17) with (16) and dividing by the trace of this product), $`{\displaystyle \frac{\overline{S}(0,t_0)e^{\beta H_0}\overline{S}(t_0,0)}{\mathrm{Tr}\left(e^{\beta H_0}\right)}}`$ $`=`$ $`{\displaystyle \frac{e^{\beta H_0}𝒫\mathrm{exp}\left(i_{i\beta }^{t_0i\beta }𝑑tH_I^{(0)}(t)\right)𝒫\mathrm{exp}\left(i_{t_0}^0𝑑tH_I^{(0)}(t)\right)}{\mathrm{Tr}\left[e^{\beta H_0}𝒫\mathrm{exp}\left(i_{i\beta }^{t_0i\beta }𝑑tH_I^{(0)}(t)\right)𝒫\mathrm{exp}\left(i_{t_0}^0𝑑tH_I^{(0)}(t)\right)\right]}}`$ (18) $`\stackrel{|t_0|\beta }{=}`$ $`{\displaystyle \frac{e^{\beta H_0}𝒫\mathrm{exp}\left(i_{i\beta }^{t_0}𝑑tH_I^{(0)}(t)\right)𝒫\mathrm{exp}\left(i_{t_0}^0𝑑tH_I^{(0)}(t)\right)}{\mathrm{Tr}\left[e^{\beta H_0}𝒫\mathrm{exp}\left(i_{i\beta }^{t_0}𝑑tH_I^{(0)}(t)\right)𝒫\mathrm{exp}\left(i_{t_0}^0𝑑tH_I^{(0)}(t)\right)\right]}}.`$ The last equality is valid perturbatively if the integration from $`t_0i\beta `$ to $`t_0`$ is negligible at every order in perturbation theory. Note that the last equality involves taking $`|t_0|`$ much greater than $`\beta `$. At zero temperature, when $`\beta \mathrm{}`$, this condition cannot hold, and since the correlation functions then may have algebraic decay with power $`1`$ at large distances, our proofs below (at equilibrium and in the steady state) do not apply. Nevertheless, as will be seen, our two-loop perturbative results for the non-equilibrium current have finite zero-temperature limit; this will be discussed further in the last section. To show the last equality in (18) we evaluate the expectation value of a local operator (or product of any local operators at fixed positions) $`𝒪`$, inserted at the right-hand side of the first equation of (18). Denoting by $$\mathrm{}_0=\frac{\mathrm{Tr}\left(e^{\beta H_0}\mathrm{}\right)}{\mathrm{Tr}\left(e^{\beta H_0}\right)}$$ (19) the averaging in the free theory at temperature $`\beta ^1`$, we consider, $$\frac{𝒫\mathrm{exp}\left(i_{i\beta }^{t_0i\beta }𝑑tH_I^{(0)}(t)\right)𝒫\mathrm{exp}\left(i_{t_0}^0𝑑tH_I^{(0)}(t)\right)𝒪_0}{𝒫\mathrm{exp}\left(i_{i\beta }^{t_0i\beta }𝑑tH_I^{(0)}(t)\right)𝒫\mathrm{exp}\left(i_{t_0}^0𝑑tH_I^{(0)}(t)\right)_0}.$$ (20) All correlation functions involved are correlation functions where the $`H_I(t)`$’s are connected to $`𝒪`$. Connected correlation functions are defined, in the usual way, by subtracting from correlation function appropriate products of expectation values. In Appendix A we recall their precise definition and main properties. Only connected correlation functions occur, because in (20) we divide by the correlation function of the operator where all the $`H_I(t)`$’s are involved. In Appendix B we show that correlation functions of the type $$H_I^{(0)}(t+t_1)H_I^{(0)}(t+t_2)\mathrm{}H_I^{(0)}(t+t_n)𝒪_0$$ factorize, as $`t\pm \mathrm{}`$, into $$H_I^{(0)}(t_1)H_I^{(0)}(t_2)\mathrm{}H_I^{(0)}(t_n)_0𝒪_0,$$ with sub-leading asymptotic contributions vanishing exponentially for finite $`\beta `$. The last step of (18) follows from the factorization property. Due to this property, connected correlation functions of the type $$H_I^{(0)}(t_1)H_I^{(0)}(t_2)\mathrm{}H_I^{(0)}(t_n)𝒪_{0,\mathrm{connected}}$$ vanish exponentially whenever any subset of consecutive time variables $`\{t_i,t_{i+1},\mathrm{},t_j\}`$ (corresponding to a subset of time-ordered operators $`H_I(t)`$’s) goes to negative infinity simultaneously. This implies that order by order in perturbation theory of (20), all integrands vanish exponentially in any large-time region, in particular in the segment $`t_0i\beta `$, which then factorizes and cancels between numerator and denominator. Hence, in the limit where $`t_0\mathrm{}`$ the last step of (18) is exact order by order in perturbation theory, and we have (15), as claimed. See Figure 2. We wish to note that our argument in Appendix B relied on the fact that $`H_I(t)`$ is a local, right-moving operator, and that it couples to the external degree of freedom $`\stackrel{}{S}`$ (the impurity) in an $`SU(2)`$-invariant way. For electronic degrees of freedom large time means large distance, and at large distances, correlation functions of local fields factorize. Combined with $`SU(2)`$ invariance, this implies the factorization of correlation functions of $`H_I(t)`$’s at large $`|t|`$. The implications of the well-defined limit $`t_0/\beta \mathrm{}`$, and in particular of the factorization at large time separation of the correlation functions involved in the perturbative coefficients, are important. It was not necessary to invoke any external relaxation mechanism: the factorization signals a decoherence in time and suggests that $`H_0`$ represents a good thermal bath, and this bath by itself provides such a mechanism. As in Caldeira-Leggett models, $`H_0`$ can be seen as an infinity of free oscillators with an appropriate frequency distribution in order to represent a thermal bath. The loss of time-reversal symmetry associated with this relaxation mechanism occurs when taking the limit $`|t_0|/\beta \mathrm{}`$. The same derivation can be carried out for more general unitary conformal field theories perturbed by an interaction $`H_I`$ at one point, or defined on a finite region of space. Inferring from our derivation, the interaction can be due to an external degree of freedom coupled to any linear combination of fields that factorize into right- and left-movers, and the coupling has to be invariant with respect to a symmetry group acting on the full configuration space. The steady state current. The derivation fails when out of equilibrium, $`V0`$. The step that becomes incorrect, if we start with expression (5) for the steady-state average, is the shifting of the integration limits $`t_0i\beta t_0`$. Indeed, the correlation functions involving $`VH_z^{(0)}(t)=VH_z`$ are not suppressed at large negative times since $`H_z`$ is a conserved charge of the Hamiltonian $`H_0`$. Moreover, due to quantum fluctuations of the charge $`H_z`$, made possible by the interaction $`H_I`$ (that is, $`[H_I,H_z]0`$), connected correlation functions involving $`H_z`$ are not zero. These two conditions are at the origin of the appearance of a non-equilibrium situation. In physical terms, the first condition is that the bath represented by $`H_0`$ does not provide a relaxation mechanism for reaching Boltzmann’s distribution of states associated to the energy $`H_0+VH_z+H_I`$; the second condition is that nevertheless, $`H_z`$ is subject to quantum fluctuations and evolves with time. In this case then the Keldysh formulation does not reduce to an equilibrium description. Our analysis, however, has not yet established that a steady state occurs. We shall present below a full proof to this effect. To motivate the proof we begin with a physical argument, based on the first formulation described around Eqs. (4) and (5), by considering the respective ground states of $`H_0`$ and $`H`$ (instead of the associated thermal density matrices) and showing that they are “far” enough and that the evolution of $`H_z`$ is “slow” enough in the limit $`L\mathrm{}`$ so that a steady state is established. Under other circumstances, we might expect some oscillating behavior. That the ground state of $`H`$ is far enough, and the evolution of $`H_z`$ is slow enough, can be made more precise in the following way. To begin with, consider the ground state $`|0`$ of $`H_0`$ and the ground state $`|V`$ of the Hamiltonian $`H_0+VH_z`$. Later we shall consider the effect of the couplings $`\lambda `$ and $`\lambda _d`$. The ground state $`|V`$ can be obtained in the following way. Consider the operator $$U_V=e^{i\frac{V}{\pi }_{\mathrm{}}^{\mathrm{}}𝑑xxJ_z(x)}$$ (21) (in order for it to be well defined, we assume an appropriate ultraviolet regularization of the operator $`J_z(x)`$). This is a unitary operator, and its effect on $`H_0`$ is: $`U_Ve^{\beta H_0}U_V`$ $`=`$ $`e^{i\frac{V}{\pi }{\scriptscriptstyle 𝑑xxJ_z(x)}}e^{\beta H_0}e^{i\frac{V}{\pi }{\scriptscriptstyle 𝑑xxJ_z(x)}}`$ (22) $`=`$ $`e^{\beta H_0}e^{i\frac{V}{\pi }{\scriptscriptstyle 𝑑xxJ_z(x+i\beta )}}e^{i\frac{V}{\pi }{\scriptscriptstyle 𝑑xxJ_z(x)}}`$ $`=`$ $`e^{\beta H_0}e^{i\frac{V}{\pi }{\scriptscriptstyle 𝑑x(xi\beta )J_z(x)}}e^{i\frac{V}{\pi }{\scriptscriptstyle 𝑑xxJ_z(x)}}`$ $`=`$ $`e^{\beta H_0}e^{\beta VH_z}e^{\frac{V^2}{2\pi ^2}{\scriptscriptstyle 𝑑x_1𝑑x_2(x_1i\beta )x_2[J_z(x_1),J_z(x_2)]}}.`$ The last exponential factor is a real number, scaling with the system size $`L`$ (it could be absorbed into the definition of $`U_V`$, at the price of losing its unitarity). Hence, the operator $`U_V`$ takes $`H_0`$ to $`H_0+VH_z`$, up to an additive number, and the ground state of $`H_0+VH_z`$ can be obtained by $$|V=U_V|0.$$ (23) Computing the expectation values of $`H_z`$ (tracing over the two-dimensional impurity space) in these ground states we have, as $`L\mathrm{}`$, $$0|H_z|0=0,V|H_z|VVL.$$ (24) In particular, $`U_V`$ has the effect, in the infinite-$`L`$ limit, of changing the asymptotic conditions to $`J_z(\pm \mathrm{})=\pi V/2`$. We discussed the ground state of $`H_0+VH_z`$. However, the actual ground state of $`H`$ yields corrections to the expectation values that are of higher order in the couplings with a finite limit as $`L\mathrm{}`$, and our conclusion therefore apply to the full Hamiltonian. While the expectation values are infinitely distant in an infinite system, the rate of change of $`H_z`$ as the interaction is switched on is finite since the operator $`H_I`$ giving rise to the current is local. This will be seen explicitly in the perturbative calculations of the current $`𝒥`$ below (recall that $`𝒥=i[H,H_z]`$). Hence, as $`L\mathrm{}`$, it would take more and more time to get from $`|0`$ to $`|V`$. Here we assume that the average of $`H_z`$ would decrease monotonically. This is expected for $`L`$ large enough and elapsed time large enough, though not infinite. More precisely in the region (6), we expect the expectation varlue of $`H_z`$ to decrease steadily; this is the steady state. In other words, we expect a steady state to occur because $`H_z`$ scales with the length of the system, whereas its variation does not. For a finite $`L`$, it does not decrease monotonically at all times, and we might eventually see an oscillating behavior of period characterized by $`L`$. We proceed now to the main result of this section: we show that to all orders in perturbation theory the limit of very large negative times, $`t_0/\beta \mathrm{}`$ in (5), is well defined for any local operator, $`𝒪`$, supported at a point or on a finite interval. This shows that there is indeed a steady state: the current operator $`𝒥`$ acquires a well-defined expectation value. Using the interaction picture with respect to $`H_0+VH_z`$, we can write the steady-state average of any operator $`𝒪`$ as $$𝒪_{ss}=\underset{t_0/\beta \mathrm{}}{lim}\frac{1}{\mathrm{Tr}\left[e^{\beta H_0}\right]}\mathrm{Tr}\left[𝒫\mathrm{exp}\left(i_0^{t_0}𝑑tH_I^{(V)}(t)\right)e^{\beta H_0}𝒫\mathrm{exp}\left(i_{t_0}^0𝑑tH_I^{(V)}(t)\right)𝒪\right]$$ (25) with $$H_I^{(V)}(t)=e^{i\left(H_0+VH_z\right)t}H_Ie^{i\left(H_0+VH_z\right)t}=e^{iVH_zt}H_I^{(0)}(t)e^{iVH_zt}.$$ (26) The operator $`H_I^{(V)}(t)`$ can be expressed in terms of “deformed” current-algebra operators. Consider $`\stackrel{}{J}_d^{(V)}(x)`$ $`=`$ $`e^{iVH_zx}\stackrel{}{J}_d(x)e^{iVH_zx}`$ $`\stackrel{}{J}_x^{(V)}(x)`$ $`=`$ $`e^{iVH_zx}\stackrel{}{J}_x(x)e^{iVH_zx}`$ $`\stackrel{}{J}_y^{(V)}(x)`$ $`=`$ $`e^{iVH_zx}\stackrel{}{J}_y(x)e^{iVH_zx}.`$ It is a simple matter to use the commutation relations (II) in order to obtain $`\stackrel{}{J}_d^{(V)}(x)`$ $`=`$ $`\stackrel{}{J}_d(x)`$ $`\stackrel{}{J}_x^{(V)}(x)`$ $`=`$ $`\mathrm{cos}(Vx)\stackrel{}{J}_x(x)+\mathrm{sin}(Vx)\stackrel{}{J}_y(x)`$ $`\stackrel{}{J}_y^{(V)}(x)`$ $`=`$ $`\mathrm{sin}(Vx)\stackrel{}{J}_x(x)+\mathrm{cos}(Vx)\stackrel{}{J}_y(x).`$ (27) Using these operators, we have $$H_I^{(V)}(t)=\lambda _d\stackrel{}{J}_d^{(V)}(t)\stackrel{}{S}+\lambda \stackrel{}{J}_y^{(V)}(t)\stackrel{}{S}.$$ (28) The proof that the limit $`t_0/\beta \mathrm{}`$ exists in (5) proceeds from arguments similar to those in the previous subsection. Let the operator $`𝒪`$ be supported on a finite interval. Moving the operator $`e^{\beta H_0}`$ to the left inside the trace in (25), we have $$𝒪_{ss}=\underset{t_0/\beta \mathrm{}}{lim}\frac{\mathrm{Tr}\left[e^{\beta H_0}𝒫\mathrm{exp}\left(i_{i\beta }^{t_0i\beta }𝑑t\stackrel{~}{H}_1^{(V)}(t)\right)𝒫\mathrm{exp}\left(i_{t_0}^0𝑑tH_I^{(V)}(t)\right)𝒪\right]}{\mathrm{Tr}\left[e^{\beta H_0}𝒫\mathrm{exp}\left(i_{i\beta }^{t_0i\beta }𝑑t\stackrel{~}{H}_I^{(V)}(t)\right)𝒫\mathrm{exp}\left(i_{t_0}^0𝑑tH_I^{(V)}(t)\right)\right]}$$ (29) where $`\stackrel{~}{H}_I^{(V)}(t)`$ $`=`$ $`e^{\beta VH_z}H_I^{(V)}(t)e^{\beta VH_z}`$ $`=`$ $`\lambda _d\stackrel{}{J}_d(t)\stackrel{}{S}+\lambda \mathrm{sin}(V(t+i\beta ))\stackrel{}{J}_x(t)\stackrel{}{S}+\lambda \mathrm{cos}(V(t+i\beta ))\stackrel{}{J}_y(t)\stackrel{}{S}.`$ The exact form of $`\stackrel{~}{H}_I`$ is actually not important; note only that it is a linear combination of local operators evolved in interaction-picture time. Again using the fact that only connected correlation functions (where $`H_I^{(V)}(t)`$ and $`\stackrel{~}{H}_I^{(V)}(t)`$ are connected to $`𝒪`$) occur order by order in perturbation theory, and the fact that correlation functions involving $`H_I^{(V)}(t)`$ and $`\stackrel{~}{H}_I^{(V)}(t)`$ factorize at large times $`t`$, one can see that all integrals are convergent in the limit $`t_0/\beta \mathrm{}`$ order by order in perturbation theory. Physically, this means that the bath represented by $`H_0`$ provides the same mechanism for the steady state to occur as the mechanism it provides for the system to reach equilibrium in the case $`V=0`$. We now cast the expression for the steady state averages in another suggestive form and derive the alternative formulation, expressed in (8), with the steady state obtained by coupling the dot (i.e. turning on the couplings $`\lambda ,\lambda _d`$) to leads initially equilibrated at temperature $`T`$ and at potential difference $`V`$. Observe that the operators with superscript $`(V)`$ form the same current algebra, Eq. (II), as those without superscript since they are obtained by the unitary transformation $`U_V`$ (21): $`\stackrel{}{J}_d^{(V)}(x)`$ $`=`$ $`U_V\stackrel{}{J}_d(x)U_V`$ $`\stackrel{}{J}_x^{(V)}(x)`$ $`=`$ $`U_V\stackrel{}{J}_x(x)U_V`$ $`\stackrel{}{J}_y^{(V)}(x)`$ $`=`$ $`U_V\stackrel{}{J}_y(x)U_V`$ $`J_z^{(V)}(x)`$ $`=`$ $`U_VJ_z(x)U_V=J_z(x)+{\displaystyle \frac{V}{2}}`$ (30) Hence, the steady-state average of an operator $`𝒪`$ can be written $$𝒪_{ss}=\underset{t_0/\beta \mathrm{}}{lim}\frac{\mathrm{Tr}\left[𝒫\mathrm{exp}\left(i_0^{t_0}𝑑tH_I^{(0)}(t)\right)U_Ve^{\beta H_0}U_V𝒫\mathrm{exp}\left(i_{t_0}^0𝑑tH_I^{(0)}(t)\right)U_V𝒪U_V\right]}{\mathrm{Tr}\left[e^{\beta H_0}\right]}$$ (31) where we recall that $`H_I^{(0)}(t)`$ is the operator evolved with $`H_0`$ only. Recalling the transformation of $`H_0`$ under $`U_V`$ (22), we find $$𝒪_{ss}=\underset{t_0/\beta \mathrm{}}{lim}\frac{\mathrm{Tr}\left[\overline{S}(0,t_0)e^{\beta \left(H_0VH_z\right)}\overline{S}(t_0,0)𝒪^{(V)}\right]}{\mathrm{Tr}\left[e^{\beta \left(H_0VH_z\right)}\right]}$$ (32) where $$𝒪^{(V)}=U_V𝒪U_V.$$ (33) In (32), it was necessary to include the factor $`e^{\beta VH_z}`$ inside the trace in the denominator of the right-hand side. For a system on a finite interval, the inclusion of this factor has the effect of cancelling the constant term that appears in (22). Then, the limit of infinite interval is well defined. Since for the current operator (12) we have $`𝒥^{(V)}=𝒥`$, this shows the equivalence, for the steady state current, between the formulation (5) and the formulation (8). In general, we will denote the steady state average in the latter formulation by $$𝒪_{ss^{}}\underset{t_0/\beta \mathrm{}}{lim}\frac{\mathrm{Tr}\left[\overline{S}(0,t_0)e^{\beta \left(H_0VH_z\right)}\overline{S}(t_0,0)𝒪\right]}{\mathrm{Tr}\left[e^{\beta \left(H_0VH_z\right)}\right]}.$$ (34) That is, $$𝒪_{ss^{}}=U_V𝒪U_V_{ss}.$$ (35) Below, we carry out some formal manipulations which are justified only if we can establish a more stringent convergence property as that used above. We need to establish that the following expression: $$𝒪_{ss^{}}=\underset{t_0/\beta \mathrm{}}{lim}\underset{t_0^{}/\beta \mathrm{}}{lim}\frac{\mathrm{Tr}\left[\overline{S}(0,t_0^{})e^{\beta \left(H_0VH_z\right)}\overline{S}(t_0,0)𝒪^{(V)}\right]}{\mathrm{Tr}\left[\overline{S}(0,t_0^{})e^{\beta \left(H_0VH_z\right)}\overline{S}(t_0,0)\right]},$$ (36) with the limits on $`t_0`$ and on $`t_0^{}`$ taken independently, will yield a result independent of the order the limits were taken. Note that we have included factors $`\overline{S}(t_0,0)`$ and $`\overline{S}(0,t_0^{})`$ in the denominator. They assure convergence and cancel by cyclicity of the trace if the limit exists. To prove convergence in (36), we consider $$𝒪_{ss}=\underset{t_0/\beta \mathrm{}}{lim}\underset{t_0^{}/\beta \mathrm{}}{lim}\frac{\mathrm{Tr}\left[𝒫\mathrm{exp}\left(i_0^{t_0^{}}𝑑tH_I^{(V)}(t)\right)e^{\beta H_0}𝒫\mathrm{exp}\left(i_{t_0}^0𝑑tH_I^{(V)}(t)\right)𝒪\right]}{\mathrm{Tr}\left[𝒫\mathrm{exp}\left(i_0^{t_0^{}}𝑑tH_I^{(V)}(t)\right)e^{\beta H_0}𝒫\mathrm{exp}\left(i_{t_0}^0𝑑tH_I^{(V)}(t)\right)\right]}.$$ (37) Indeed, the same arguments we used to establish the connectedness and factorization allow us to take, for instance, first the limit with $`|t_0^{}|`$ large, then the limit with $`|t_0|`$ large, or vice versa. The result is unique. Using the operator $`U_V`$ in a manner similar to the one above, it is a simple matter to obtain (36) from (37) and the result then is the same as the one obtained from the formulation (34). Alternative description of the Steady State. What replaces the density matrix $`e^{\beta H}`$ description of equilibrium? We could translate the proof establishing equilibrium when ($`V=0`$) to the proof establishing steady state when ($`V0`$) by means of the current algebra of symmetries. By similar means we shall show that a new operator will play for the system in its steady state a similar role to the one played by the density matrix in equilibrium. Such a steady-state density matrix can be obtained from simple manipulations, now that we have established the convergence of the integrals. Consider the formulation (36) of the steady-state problem, with $`𝒪`$ in (33) an operator supported on a finite interval in the theory $`H_0`$. Bringing $`e^{\beta H_0}`$ completely to the left, the right-hand side of (36) can be written as follows: $$\underset{t_0/\beta \mathrm{}}{lim}\underset{t_0^{}/\beta \mathrm{}}{lim}\frac{\mathrm{Tr}\left[e^{\beta H_0}𝒫\mathrm{exp}\left(i_{i\beta }^{t_0^{}i\beta }𝑑tH_I^{(0)}(t)\right)e^{\beta VH_z}𝒫\mathrm{exp}\left(i_{t_0}^0𝑑tH_I^{(0)}(t)\right)𝒪^{(V)}\right]}{\mathrm{Tr}\left[e^{\beta H_0}𝒫\mathrm{exp}\left(i_{i\beta }^{t_0^{}i\beta }𝑑tH_I^{(0)}(t)\right)e^{\beta VH_z}𝒫\mathrm{exp}\left(i_{t_0}^0𝑑tH_I^{(0)}(t)\right)\right]}.$$ Since we showed that the limits can be taken independently, we can shift $`t_0^{}i\beta `$ to $`t_0^{}`$ both in the numerator and in the denominator, without shifting $`t_0`$, with vanishing error in the limit. We can then take $`t_0^{}=t_0`$ and keep only one limit symbol. Inserting $$1=𝒫\mathrm{exp}\left(i_{t_0}^0𝑑tH_I^{(0)}(t)\right)𝒫\mathrm{exp}\left(i_0^{t_0}𝑑tH_I^{(0)}(t)\right)$$ just before the operator $`e^{\beta VH_z}`$ we have, $$\underset{t_0/\beta \mathrm{}}{lim}\frac{\mathrm{Tr}\left[e^{\beta H|_{V=0}}𝒫\mathrm{exp}\left(i_0^{t_0}𝑑tH_I^{(0)}(t)\right)e^{\beta VH_z}𝒫\mathrm{exp}\left(i_{t_0}^0𝑑tH_I^{(0)}(t)\right)𝒪^{(V)}\right]}{\mathrm{Tr}\left[e^{\beta H|_{V=0}}𝒫\mathrm{exp}\left(i_0^{t_0}𝑑tH_I^{(0)}(t)\right)e^{\beta VH_z}𝒫\mathrm{exp}\left(i_{t_0}^0𝑑tH_I^{(0)}(t)\right)\right]}.$$ Defining the operator: $$Y=\underset{t_0/\beta \mathrm{}}{lim}\overline{S}(0,t_0)H_z\overline{S}(t_0,0)$$ (38) allows us to write the steady-state average of a local operator as $$𝒪_{ss^{}}=\frac{\mathrm{Tr}\left[e^{\beta H|_{V=0}}e^{\beta VY}𝒪\right]}{\mathrm{Tr}\left[e^{\beta H|_{V=0}}e^{\beta VY}\right]}.$$ (39) Note that the limit (38) cannot be expected to exist as an operator (recall that we are dealing with operators with infinitely many matrix elements), but only when inserted into appropriate traces (or only when appropriate matrix elements are considered). More precisely, we have only proven that (38) is a well-defined operator when it is evaluated in expressions like (39), and that the result is the steady-state average of the local operators inserted. This is a statement solely about a small part of the matrix elements of the operator (38). The meaning of Eq. (39) is that one must first evaluate the traces and their ratio with the expression (38) at finite $`t_0/\beta `$, then take the limit indicated in (38) on the result. The properties of $`Y`$ as an operator acting in a Hilbert space will be discussed elsewhere. Observe, however, that in all situations where the operator $`Y`$ is well defined (that is, when we consider the appropriate matrix elements), then it is a conserved charge. Indeed, when it is well defined, then the limit $`t_0/\beta \mathrm{}`$ of $`S(0,t_0)H_zS(t_0,0)`$ (or of any function of this operator) must exist. Since $`S(t_1,t_2)=S(t_1+dt,t_2+dt)`$, we have $`\frac{d}{dt}lim_{t_0/\beta \mathrm{}}S(t,t_0)H_zS(t_0,t)=0`$, hence $`[H|_{V=0},Y]=0`$. Then we can finally write $$𝒪_{ss^{}}=\frac{\mathrm{Tr}\left[e^{\beta (H|_{V=0}VY)}𝒪\right]}{\mathrm{Tr}\left[e^{\beta (H|_{V=0}VY)}\right]}.$$ (40) That is, averages in the steady-state can be obtained by tracing with an appropriate density matrix<sup>2</sup><sup>2</sup>2Observe here that the operator $`H|_{V=0}VY`$ need not have a spectrum bounded from below in the Hilbert space of $`H|_{V=0}`$. As usual, in any explicit evaluation, the traces are regularized using a regularisation that respects cyclicity. The resulting expression is well defined with insertion of local operators only and the limit implied in the definition of the operator $`Y`$ should be taken simultaneously in the numerator and the denominator.. What difference is there between the equilibrium and the steady state? Consider a quantum mechanical system described by a Hamiltonian $`H`$. Put the system at equilibrium with a bath where there can be exchange of heat and of any quantity $`Q`$ that is conserved by the dynamics $`H`$. The average of observables is then described by the density matrix $`e^{\beta (H+\mu Q)}`$ where $`\mu `$ is the chemical potential associated to $`Q`$: the energy brought to the system by increasing $`Q`$ by one unit. In expression (40), the average of a local operator $`𝒪`$ in the steady state is a trace with a density matrix of exactly the same form. The main difference is that the operator $`Y`$ is a non-local conserved charge. A local conserved charge can be written as an integral over space of a local operator of the theory $`H|_{V=0}`$ plus a local operator at the impurity site, with possible non-trivial impurity-space components. The operator $`Y`$ (38) cannot be written in that way. To see this, we can write it as follows: $$Y=H_z+_{\mathrm{}}^0𝑑t𝒥(t)$$ (41) where $$𝒥(t)=\overline{S}(0,t)𝒥\overline{S}(t,0)$$ (42) is the time-evolved current $`𝒥`$ with respect to the theory $`H|_{V=0}`$. Then it is simple to write it as an integral of a charge density: $$Y=_{\mathrm{}}^{\mathrm{}}𝑑xj^{tot}(x,0)$$ (43) with $$j^{tot}(x,t)=\frac{1}{\pi }J_z(x,t)+\delta (x)_{\mathrm{}}^t𝑑t^{}𝒥(t^{}).$$ (44) The charge density has a local bulk part, but the term at the impurity is not a local field of the theory $`H|_{V=0}`$: it is the time integral of the current, and the current is not the time derivative of a local field. The non-locality of $`Y`$ is the main difference between the description of a steady state and of an equilibrium state. In the formulation (40), only a restricted set of operators $`𝒪`$ have well-defined average: those that have stationary expectation values. All local operators are of this type, but, for instance, it is simple to see that the operator $`H_z`$ does not have a well-defined steady-state value. Note also that the operator $`Y`$ gives in principle a description of the asymptotic state that one can use in order to describe the steady state: quantities in the steady state can be evaluated as averages in an appropriate asymptotic state. Further analysis in this direction will be presented in our future works. We wish to remark that sometime ago Hershfield Hershfield93 has considered steady-state flow and has argued that under some assumptions concerning the relaxation of correlation function an expression (40) would govern the steady-state current. He gave then implicit equations to determine $`Y`$. It appears to us that our explicit expressions for the operator $`Y`$ satisfies his implicit equations, and should hence correspond to the same operator (although we have not thoroughly ascertained the confluence of the two approaches). We must stress, however, that no assumptions were made in our derivation. ## IV RG-improved real-time perturbation theory The perturbative expansion. We now turn to real-time perturbation theory for the current (8). We take the formulation where the system is initially brought to equilibrium with a nonzero bias voltage, then disconnected from the external bath before the voltage is turned off and the interaction is turned on. It will be convenient to consider adiabatically turning on the interaction in the infinite past: we introduce a large-time exponential cutoff, $`e^{\eta t}H_I`$, with $`\eta `$ a positive scale with dimension of energy, and take the limit $`t_0/\beta \mathrm{}`$ in (8). The quantity $`\eta \beta `$ will be sent to 0 at the end of the calculations. This means, physically, that the two leads are slowly brought towards the dot after the voltage has been turned off. Our proof that there are no divergencies as $`t_0/\beta \mathrm{}`$ in the previous section shows that there are no divergencies as $`\eta \beta 0`$. The current can then be written $$𝒥_{ss}=\underset{\eta \beta 0^+}{lim}\frac{\mathrm{Tr}\left[e^{\beta (H_0VH_z)}S_\eta (\mathrm{},0)𝒥S_\eta (0,\mathrm{})\right]}{\mathrm{Tr}\left[e^{\beta (H_0VH_z)}\right]}$$ (45) with <sup>3</sup><sup>3</sup>3By definition, $`dS_\eta (t_1,t_2)/dt_2=iS_\eta (t_1,t_2)(H_0+e^{\eta t_2}H_I)`$ and $`S_\eta (t,t)=1`$. $$S_\eta (t_1,t_2)=𝒫\mathrm{exp}_{t_1}^{t_2}i\left(H_0+e^{\eta t}H_I\right)𝑑t.$$ (46) More precisely, $$𝒥_{ss}=\underset{\eta \beta 0^+}{lim}\underset{k=0}{\overset{\mathrm{}}{}}i^k_{\mathrm{}}^0𝑑t_1e^{\eta t_1}_{t_1}^0𝑑t_2e^{\eta t_2}\mathrm{}_{t_{k1}}^0𝑑t_ke^{\eta t_k}[H_I(t_1),[H_I(t_2),\mathrm{},[H_I(t_k),𝒥]\mathrm{}]]_V$$ (47) where $$\mathrm{}_V=\frac{\mathrm{Tr}\left[e^{\beta \left(H_0VH_z\right)}\mathrm{}\right]}{\mathrm{Tr}\left[e^{\beta \left(H_0VH_z\right)}\right]}.$$ (48) The integrals in this expansion are plagued with ultraviolet divergencies which we have to regularize. This can be done in several ways. For our purposes, it will be convenient to modify the operators $`\stackrel{}{J}_d(x),\stackrel{}{J}_x(x),\stackrel{}{J}_y(x)`$ in order to render their correlation functions regular at coinciding points. More precisely, we choose the regularization scheme where all operators (in the Hamiltonian and in correlation functions) at the impurity site are regularized, whereas all operators away from it are unaffected. Since the interaction is only at the impurity site, this is enough to regularize the theory. Define the momentum space (mode) operators $`\stackrel{}{J}_d(p),\stackrel{}{J}_x(p),\stackrel{}{J}_y(p)`$ and $`J_z(p)`$ in the following way: $`\stackrel{}{J}_d(x)={\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑p\stackrel{}{J}_d(p)e^{ipx},\stackrel{}{J}_x(x)={\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑p\stackrel{}{J}_x(p)e^{ipx},`$ $`\stackrel{}{J}_y(x)={\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑p\stackrel{}{J}_y(p)e^{ipx},J_z(x)={\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑pJ_z(p)e^{ipx}.`$ The mode operators satisfy the following set of commutation relations: $`[J_d^i(p),J_d^j(q)]`$ $`=`$ $`iϵ_{ijk}J_d^k(p+q)+p\delta _{ij}\delta (p+q)`$ $`[J_x^i(p),J_x^j(q)]`$ $`=`$ $`iϵ_{ijk}J_d^k(p+q)+p\delta _{ij}\delta (p+q)`$ $`[J_y^i(p),J_y^j(q)]`$ $`=`$ $`iϵ_{ijk}J_d^k(p+q)+p\delta _{ij}\delta (p+q)`$ $`[J_z(p),J_z(q)]`$ $`=`$ $`p\delta (p+q)`$ $`[J_d^i(p),J_x^j(q)]`$ $`=`$ $`iϵ_{ijk}J_x^k(p+q)`$ (49) $`[J_d^i(p),J_y^j(q)]`$ $`=`$ $`iϵ_{ijk}J_y^k(p+q)`$ $`[J_d^i(p),J_z(q)]`$ $`=`$ $`0`$ $`[J_x^i(p),J_y^j(q)]`$ $`=`$ $`i\delta _{ij}J_z(p+q)`$ $`[J_y^i(p),J_z(q)]`$ $`=`$ $`iJ_x^i(p+q)`$ $`[J_z(p),J_x^i(q)]`$ $`=`$ $`iJ_y^i(p+q).`$ We then introduce the regularized operators $`(\stackrel{}{J}_d)_\mathrm{\Lambda }(x)={\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑pR_\mathrm{\Lambda }(p)\stackrel{}{J}_d(p)e^{ipx},(\stackrel{}{J}_x)_\mathrm{\Lambda }(x)={\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑pR_\mathrm{\Lambda }(p)\stackrel{}{J}_x(p)e^{ipx},`$ $`(\stackrel{}{J}_y)_\mathrm{\Lambda }(x)={\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑pR_\mathrm{\Lambda }(p)\stackrel{}{J}_y(p)e^{ipx}`$ where $`R_\mathrm{\Lambda }(p)`$ is a function that vanishes as $`|p|\mathrm{}`$, and whose complex conjugate satisfies $`R_\mathrm{\Lambda }(p)^{}=R_\mathrm{\Lambda }(p)`$ in order to preserve hermiticity of the regularized operators. The function $`R_\mathrm{\Lambda }(p)`$ can be chosen in many ways, and the choice is a matter of convenience. We will choose a gaussian regularization, $$R_\mathrm{\Lambda }(p)=e^{p^2/(2\mathrm{\Lambda }^2)}.$$ (50) The parameter $`\mathrm{\Lambda }`$ plays the role of an effective band width (we do not denote it $`D`$ since it is certainly not exactly the band width). The universal part of the limit $`\mathrm{\Lambda }V,T`$ is the same as that of the limit of large band width. The time integrals in (47) can now be traded to momentum integrals: $`𝒥_{ss}`$ $`=`$ $`{\displaystyle \underset{k=0}{\overset{\mathrm{}}{}}}(1)^k{\displaystyle \frac{dp_1R_\mathrm{\Lambda }(p_1)}{p_1+i\eta }\frac{dp_2R_\mathrm{\Lambda }(p_2)}{p_1+p_2+2i\eta }\mathrm{}\frac{dp_kR_\mathrm{\Lambda }(p_k)}{p_1+p_2+\mathrm{}+p_k+ki\eta }}`$ (51) $`{\displaystyle 𝑑sR_\mathrm{\Lambda }(s)[\stackrel{~}{H}_I(p_1),[\stackrel{~}{H}_I(p_2),\mathrm{},[\stackrel{~}{H}_I(p_k),\stackrel{~}{𝒥}(s)]\mathrm{}]]_V}`$ where $$\stackrel{~}{H}_I(p)=\lambda _d\stackrel{}{J}_d(p)\stackrel{}{S}+2\stackrel{}{J}_y(p)\stackrel{}{S}$$ (52) and $$\stackrel{~}{𝒥}(p)=\stackrel{}{J}_x(p)S.$$ (53) The parameter $`\eta `$ can be set to zero with the requirement that the momentum integrals be taken on a line parallel to the real axis in the $`p`$-plane with a slight positive imaginary part: $`𝒥_{ss}`$ $`=`$ $`{\displaystyle \underset{k=0}{\overset{\mathrm{}}{}}}(1)^k{\displaystyle _+}{\displaystyle \frac{dp_1R_\mathrm{\Lambda }(p_1)}{p_1}}{\displaystyle _+}{\displaystyle \frac{dp_2R_\mathrm{\Lambda }(p_2)}{p_1+p_2}}\mathrm{}{\displaystyle _+}{\displaystyle \frac{dp_kR_\mathrm{\Lambda }(p_k)}{p_1+p_2+\mathrm{}+p_k}}`$ (54) $`{\displaystyle 𝑑qR_\mathrm{\Lambda }(q)[\stackrel{~}{H}_I(p_1),[\stackrel{~}{H}_I(p_2),\mathrm{},[\stackrel{~}{H}_I(p_k),\stackrel{~}{𝒥}(q)]\mathrm{}]]_V}.`$ Note that this can be done since $`R_\mathrm{\Lambda }(p)`$ does not have a singularity at $`p=0`$ or anywhere on the real $`p`$-line. No singularity at $`p_1+\mathrm{}+p_j=0`$ (for $`j=1,\mathrm{},k`$) can occur in the averages $`\mathrm{}_V`$ in the expression above, since we know that there are no divergencies as $`\eta \beta 0`$. In fact, for explicit calculations, it will be more convenient to integrate over the real line in momentum space, and to use the formalism of principal value integrals. For this purpose, recall that $$_+𝑑p\frac{f(p)}{(p+q)^n}=𝑑p\left(\frac{(1)^n}{(n1)!}i\pi \delta ^{(n1)}(p+q)+\underset{¯}{\mathrm{P}}\frac{1}{(p+q)^n}\right)f(p)$$ (55) where $`\underset{¯}{\mathrm{P}}`$ says that we must take the principal value of the integral: $$𝑑p\left(\underset{¯}{\mathrm{P}}\frac{1}{(p+q)^n}\right)f(p)=\text{finite part of}\left(_{\mathrm{}}^ϵ+_ϵ^{\mathrm{}}\right)dp\frac{f(p)}{(p+q)^n}\text{in power expansion as }ϵ0.$$ (56) Explicit one-loop calculations. The trace over the impurity space of the operators $$[\stackrel{~}{H}_I(p_1),[\stackrel{~}{H}_I(p_2),\mathrm{},[\stackrel{~}{H}_I(p_k),\stackrel{~}{𝒥}(q)]\mathrm{}]]$$ can be obtained by using the following general formula, valid for any vector operators $`\stackrel{}{A}`$ and $`\stackrel{}{B}`$ that commute with $`\stackrel{}{S}`$: $$[\stackrel{}{A}\stackrel{}{S},\stackrel{}{B}\stackrel{}{S}]=\frac{i}{2}\{A^i,B^j\}ϵ_{ijk}S^k+\frac{1}{4}[A^i,B^i]$$ (57) where $`\{,\}`$ is the anti-commutator. Using the commutation relations (IV) for evaluating the commutators, one is left then only with multiple anti-commutators of the mode operators. To one loop we will need the following commutators: $`C_1`$ $`=`$ $`[\stackrel{}{J}_y(q)\stackrel{}{S},\stackrel{}{J}_x(k)\stackrel{}{S}]`$ $`C_2`$ $`=`$ $`[\stackrel{}{J}_d(p)\stackrel{}{S},[\stackrel{}{J}_y(q)\stackrel{}{S},\stackrel{}{J}_x(k)\stackrel{}{S}]]`$ $`C_3`$ $`=`$ $`[\stackrel{}{J}_y(p)\stackrel{}{S},[\stackrel{}{J}_d(q)\stackrel{}{S},\stackrel{}{J}_x(k)\stackrel{}{S}]].`$ (58) Under trace over the impurity space, they give $`\mathrm{Tr}_{\mathrm{impurity}}(C_1)`$ $`=`$ $`3iJ_z(q+k)`$ $`\mathrm{Tr}_{\mathrm{impurity}}(C_2)`$ $`=`$ $`2\{J_y^i(q),J_x^i(p+k)\}2\{J_y^i(p+q),J_x^i(k)\}`$ $`\mathrm{Tr}_{\mathrm{impurity}}(C_3)`$ $`=`$ $`2\{J_y^i(p+q),J_x^i(k)\}.`$ (59) The traces $`\mathrm{}_V`$ over the bulk-CFT Hilbert space of these anti-commutators can be calculated without the need to construct the Hilbert space for $`H_0VH_z`$, but only by using the following exchange relations: $`\stackrel{}{J}_d(p)e^{\beta \left(H_0VH_z\right)}`$ $`=`$ $`e^{\beta \left(H_0VH_z\right)}\stackrel{}{J}_d(p)e^{\beta p}`$ $`J_z(p)e^{\beta \left(H_0VH_z\right)}`$ $`=`$ $`e^{\beta \left(H_0VH_z\right)}J_z(p)e^{\beta p}`$ $`\stackrel{}{J}_+(p)e^{\beta \left(H_0VH_z\right)}`$ $`=`$ $`e^{\beta \left(H_0VH_z\right)}\stackrel{}{J}_+(p)e^{\beta (p+V)}`$ $`\stackrel{}{J}_{}(p)e^{\beta \left(H_0VH_z\right)}`$ $`=`$ $`e^{\beta \left(H_0VH_z\right)}\stackrel{}{J}_{}(p)e^{\beta (pV)}`$ (60) where we use the following linear combinations: $$\stackrel{}{J}_+(p)=\frac{1}{2}\left(\stackrel{}{J}_x(p)+i\stackrel{}{J}_y(p)\right),\stackrel{}{J}_{}(p)=\frac{1}{2}\left(\stackrel{}{J}_x(p)i\stackrel{}{J}_y(p)\right).$$ (61) The exchange relations imply for any operator $`𝒪`$ $`𝒪\stackrel{}{J}_d(p)_V`$ $`=`$ $`e^{\beta p}\stackrel{}{J}_d(p)𝒪_V`$ $`𝒪J_z(p)_V`$ $`=`$ $`e^{\beta p}J_z(p)𝒪_V`$ $`𝒪\stackrel{}{J}_+(p)_V`$ $`=`$ $`e^{\beta (p+V)}\stackrel{}{J}_+(p)𝒪_V`$ $`𝒪\stackrel{}{J}_{}(p)_V`$ $`=`$ $`e^{\beta (pV)}\stackrel{}{J}_{}(p)𝒪_V.`$ (62) Using the commutator formulas (IV), using these relations and using the traces of single operators, all traces can be calculated. Traces of single operators can be calculated from the fact that $`\stackrel{}{J}_{d,x,y}_0=J_z_0=0`$, and from $`𝒪_V=U_V𝒪U_V_0`$. In terms of modes, the operator $`U_V`$ has the representation $$U_V=e^{2VJ_z^{}(0)}$$ (63) where we formally define $`J_z^{}(p)=dJ_z(p)/dp`$. The traces of single operators are then given by $$\stackrel{}{J}_{d,x,y}(p)_V=0,J_z(p)_V=V\delta (p).$$ (64) The trace of the only type of anti-commutator appearing in (59) is then given by $$\{J_y^i(p),J_x^j(q)\}_V=\frac{i}{8}F(p)\delta (p+q)\delta _{ij}$$ (65) where $$F(p)=(p+V)\frac{1+e^{\beta (p+V)}}{1e^{\beta (p+V)}}(pV)\frac{1+e^{\beta (pV)}}{1e^{\beta (pV)}}.$$ (66) From (54) and using (55), the integrals to be calculated at one loop are $`I_1`$ $`=`$ $`{\displaystyle 𝑑qR_\mathrm{\Lambda }(q)\left(i\pi \delta (q)+\underset{¯}{\mathrm{P}}\frac{1}{q}\right)𝑑kR_\mathrm{\Lambda }(k)C_1_V}`$ $`I_{2,3}`$ $`=`$ $`{\displaystyle }dpR_\mathrm{\Lambda }(p)(i\pi \delta (p)+\underset{¯}{\mathrm{P}}{\displaystyle \frac{1}{p}}){\displaystyle }dqR_\mathrm{\Lambda }(q)(i\pi \delta (p+q)+\underset{¯}{\mathrm{P}}{\displaystyle \frac{1}{p+q}})\times `$ (67) $`{\displaystyle 𝑑kR_\mathrm{\Lambda }(k)C_{2,3}_V}.`$ The current will then be given by $$𝒥_{ss}=\lambda ^2\left(I_1+\lambda _d(I_2+I_3)+O(\lambda _d^2,\lambda ^2)\right).$$ (68) Since all integrals are real, it is clear that in the expressions (67), only terms with an odd number of delta functions in the momentum variables give non-zero contributions. It is a simple matter then to obtain $`I_1`$ $`=`$ $`{\displaystyle \frac{3\pi V}{2}}`$ (69) $`I_2=I_3`$ $`=`$ $`{\displaystyle \frac{3\pi V}{4}}{\displaystyle 𝑑p\underset{¯}{\mathrm{P}}\frac{1}{p}f(p)R_\mathrm{\Lambda }(Vp)^2}`$ (70) where $`p`$ is now a dimensionless momentum variable, and $$f(p)=(p+1)\frac{1+e^{w(p+1)}}{1e^{w(p+1)}}(p1)\frac{1+e^{w(p1)}}{1e^{w(p1)}}$$ (71) and $$w=\beta V.$$ (72) Using the symmetry $`f(p)=f(p)`$, the integral $`I_2`$ can be calculated in the following way, keeping only the divergent and finite parts as $`\mathrm{\Lambda }\mathrm{}`$: $`I_2=I_3`$ $`=`$ $`{\displaystyle \frac{3\pi V}{2}}{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dp}{p}}f(p)e^{p^2V^2/\mathrm{\Lambda }^2}`$ (73) $``$ $`{\displaystyle \frac{3\pi V}{2}}\left[{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dp}{p}}(f(p)2+2e^{p^2})2{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dp}{p}}(e^{p^2}1)e^{p^2V^2/\mathrm{\Lambda }^2}\right]`$ $``$ $`3\pi V\left[P(w)+\mathrm{ln}\left({\displaystyle \frac{\mathrm{\Lambda }}{V}}\right)\right]`$ where the symbol $``$ means that equality is valid only for the finite and divergent parts as $`\mathrm{\Lambda }\mathrm{}`$ and where $$P(w)=\frac{1}{2}_0^{\mathrm{}}\frac{dp}{p}(f(p)2+2e^{p^2})=_0^{\mathrm{}}\frac{dp}{p}\left(\frac{p+1}{e^{w(p+1)}1}\frac{p1}{e^{w(p1)}1}+e^{p^2}\right).$$ (74) This is a well-defined function for all $`w`$ with positive real part, and is easy to evaluate numerically. It will be useful to have the asymptotic behavior of the function $`P(w)`$, at large and small $`w`$. This is evaluated in Appendix D. The asymptotic expansion at large $`w`$ is given by $$P(w)1+\frac{\gamma }{2}\frac{\pi ^2}{3}w^2\text{as}w\mathrm{},$$ (75) whereas the expansion at small $`w`$ is $$P(w)=\mathrm{ln}(w)+\frac{\sqrt{\pi }}{2}+\gamma \mathrm{ln}(2\pi )+O(w)\text{as}w0.$$ (76) Summarizing, the zero- and one-loop divergent and finite contributions to the current are (with $`T=1/\beta `$) $$𝒥_{ss}=3\pi V\lambda ^2\left[\frac{1}{2}+2\lambda _d\left(P\left(\frac{V}{T}\right)+\mathrm{ln}\left(\frac{\mathrm{\Lambda }}{V}\right)\right)\right]$$ (77) with $`P(w)`$ given by the integrals in (74), $`P(w)=_0^{\mathrm{}}\frac{dp}{p}\left(\frac{p+1}{e^{w(p+1)}1}\frac{p1}{e^{w(p1)}1}+e^{p^2}\right)`$. Note that only the combination $`P(V/T)+\mathrm{ln}(\mathrm{\Lambda }/V)`$ appears at one loop. This combination has the limits $`P\left({\displaystyle \frac{V}{T}}\right)+\mathrm{ln}\left({\displaystyle \frac{\mathrm{\Lambda }}{V}}\right)`$ $``$ $`\left(\mathrm{ln}\left({\displaystyle \frac{V}{\mathrm{\Lambda }}}\right)1+{\displaystyle \frac{\gamma }{2}}\right)\text{as}TV\mathrm{\Lambda }`$ (78) $``$ $`\left(\mathrm{ln}\left({\displaystyle \frac{T}{\mathrm{\Lambda }}}\right){\displaystyle \frac{\sqrt{\pi }}{2}}\gamma +\mathrm{ln}(2\pi )\right)\text{as}VT\mathrm{\Lambda }.`$ The one-loop calculation was also performed in KaminskiNG00 (although the analysis did not go in as much detail as ours), and it can be verified that their results agree with ours. Results for the two-loop calculations. Two-loop integrals come from the commutators (here $`p`$, $`q`$, $`r`$ and $`k`$ are all momentum variables): $`C_4`$ $`=`$ $`[\stackrel{}{J}_d(p)\stackrel{}{S},[\stackrel{}{J}_d(q)\stackrel{}{S},[\stackrel{}{J}_y(r)\stackrel{}{S},\stackrel{}{J}_x(k)\stackrel{}{S}]]]`$ $`C_5`$ $`=`$ $`[\stackrel{}{J}_d(p)\stackrel{}{S},[\stackrel{}{J}_y(q)\stackrel{}{S},[\stackrel{}{J}_d(r)\stackrel{}{S},\stackrel{}{J}_x(k)\stackrel{}{S}]]]`$ $`C_6`$ $`=`$ $`[\stackrel{}{J}_y(p)\stackrel{}{S},[\stackrel{}{J}_d(q)\stackrel{}{S},[\stackrel{}{J}_d(r)\stackrel{}{S},\stackrel{}{J}_x(k)\stackrel{}{S}]]]`$ $`C_7`$ $`=`$ $`[\stackrel{}{J}_y(p)\stackrel{}{S},[\stackrel{}{J}_y(q)\stackrel{}{S},[\stackrel{}{J}_y(r)\stackrel{}{S},\stackrel{}{J}_x(k)\stackrel{}{S}]]]`$ (79) appearing inside the traces in the integrands of (54). Calculating the traces and using (55), we can find the corresponding set of two-loop integrals, $`I_4,I_5,I_6`$ and $`I_7`$ (written in Appendix E). These integrals enter the current as $$𝒥_{ss}=\lambda ^2\left(I_1+\lambda _d(I_2+I_3)+\lambda _d^2(I_4+I_5+I_6)+\lambda ^2I_7+O(\lambda _d^3,\lambda ^2\lambda _d)\right).$$ (80) Explicit calculation of these integrals lead to the following divergent parts as $`\mathrm{\Lambda }\mathrm{}`$: $`I_4+I_5+I_6`$ $``$ $`3\pi V\left[10P(w)+5\mathrm{ln}\left({\displaystyle \frac{\mathrm{\Lambda }}{V}}\right)1\right]\mathrm{ln}\left({\displaystyle \frac{\mathrm{\Lambda }}{V}}\right)+\mathrm{finite}`$ $`I_7`$ $``$ $`3\pi V\left[2P(w)+\mathrm{ln}\left({\displaystyle \frac{\mathrm{\Lambda }}{V}}\right)1\right]\mathrm{ln}\left({\displaystyle \frac{\mathrm{\Lambda }}{V}}\right)+\mathrm{finite}.`$ (81) The finite contributions are much more complicated, and are reported in Appendix F. Let us only notice that the limit $`T/V0`$ of these contributions is finite, as in the one-loop results. This indicates that the voltage $`V`$ plays the role of a good infrared cutoff for the perturbative calculation of the current to two loops: the temperature may be set to zero without divergences to this order. Summarizing, the zero-, one- and two-loop finite and divergent contributions to the current are, $`𝒥_{ss}`$ $`=`$ $`3\pi V\lambda ^2\{{\displaystyle \frac{1}{2}}+2\lambda _d(P(w)+\mathrm{ln}\left({\displaystyle \frac{\mathrm{\Lambda }}{V}}\right))+`$ (82) $`+\lambda _d^2[10P(w)+5\mathrm{ln}\left({\displaystyle \frac{\mathrm{\Lambda }}{V}}\right)1]\mathrm{ln}\left({\displaystyle \frac{\mathrm{\Lambda }}{V}}\right)+\lambda ^2[2P(w)+\mathrm{ln}\left({\displaystyle \frac{\mathrm{\Lambda }}{V}}\right)1]\mathrm{ln}\left({\displaystyle \frac{\mathrm{\Lambda }}{V}}\right)\}+`$ $`+\lambda ^2\lambda _d^2[I_4+I_5+I_6]_{\mathrm{finite}}+\lambda ^4[I_7]_{\mathrm{finite}}`$ where $`[I_4+I_5+I_6]_{\mathrm{finite}}`$ and $`[I_7]_{\mathrm{finite}}`$ are given in Appendix F. The Renormalization Group equation (The Callan-Symanzik equation). In systems at equilibrium, Wilson’s renormalization group ideas allow us to understand how physical quantities can have universal forms (independent of the precise form of the interactions at the microscopic level) when all physical energy scales (temperature, voltage, etc.) are much lower than the microscopic scales (band width, inverse lattice spacing, etc.). Out of equilibrium, it is not obvious that Wilson’s renormalization group ideas still apply. A different, but equivalent, way to look at universality is to study the limit where the cut-off $`\mathrm{\Lambda }`$ is increased and sent to infinity, while the couplings dependence on the cut-off is again governed by an RG equation, valid at very large cut-off. If such a limit exists then all quantities tend to their universal form. We will argue that the steady-state average of the current operator (or of any operator having a well-defined average in the steady state) satisfies the Callan-Symanzik equation with the same beta function and anomalous dimension as they occur in any average evaluated at equilibrium. More precisely, we will argue that $$\left(\mathrm{\Lambda }\frac{}{\mathrm{\Lambda }}|_{\lambda ,\lambda _d}+\beta _\lambda (\lambda ,\lambda _d)\frac{}{\lambda }+\beta _{\lambda _d}(\lambda ,\lambda _d)\frac{}{\lambda _d}\right)𝒥_{ss}\stackrel{\mathrm{\Lambda }\mathrm{}}{=}0$$ (83) where $`\beta _\lambda `$ and $`\beta _{\lambda _d}`$ are the beta functions of the anisotropic two-channel Kondo model. Note that the anomalous dimension term does not occur: the current operator $`𝒥`$ has zero anomalous dimension (this is natural from a physical perspective, as the current is a physical object which should not change with a change of scales; we will verify this explicitly to one loop, and indirectly to two loops, below). We should note that the Callan-Symanzik equations with one-loop beta functions and zero anomalous dimension was written in KaminskiNG00 for the steady-state average of the current, from physical arguments<sup>4</sup><sup>4</sup>4The authors also considered the case of an oscillating applied voltage, where they had to introduce an extra scale, a “decoherence time,” in order to recover universality. We do not consider this case here.. Here we present a quantum field theoretic argument that applies to all orders (and all matrix elements), and in the next section we explicitly verify this argument and calculate the beta functions and anomalous dimension in a universal fashion (so that it automatically applies to the steady state) to one-loop order. The Callan-Symanzik equation embodies Wilson’s renormalization group ideas: it tells us how a change of cutoff $`\mathrm{\Lambda }`$ (for instance, the band width) can be compensated by a change of few relevant coupling constant, as long as all physical energy scales are much lower than $`\mathrm{\Lambda }`$. Solving the Callan-Symanzik equations (this is done below for the current in the steady state) allows us to describe the low energy behavior of the steady-state current in terms of the ratios $`V/T_K`$ and $`T/T_K`$, where $`T_K`$ is an integration constant, as well as of one extra parameter (invariant under the RG flow) characterizing the asymmetry between the couplings $`\lambda _d`$ and $`\lambda `$; we will denote this parameter by $`C`$. The integration constant $`T_K`$ and the extra parameter $`C`$ characterize the quantum field theory; when they are fixed, all averages can be evaluated unambiguously. These parameters are not universal: different microscopic theories have low-energy behaviors described by different values for them. Up to these non-universal quantities, the quantum field theory description is universal, independent of the precise choice of the cutoff procedure (precise structure of the band, for instance) and of irrelevant couplings (interactions that give vanishing contributions at low energies). The integration constant $`T_K`$ is the Kondo temperature: the temperature above which the “Kondo cloud” gets destroyed by the thermal energy. It is related to the microscopic values of the couplings $`\lambda `$ and $`\lambda _d`$ (the values when $`\mathrm{\Lambda }`$ is chosen to be of the order of the real band width), and it decreases if the couplings are decreased. At zero couplings, the Kondo temperature is zero and all scales of the low-energy physics disappear: this is a quantum critical point. The quantum field theory with finite ratios $`V/T_K`$ and $`T/T_K`$ describes the situation $`T_K\mathrm{\Lambda }`$: the couplings are sent to zero at the same time as the voltage and temperature are made much smaller than $`\mathrm{\Lambda }`$. This is the scaling limit, describing the region around the quantum critical point $`\lambda =\lambda _d=0`$. In particular, one finds that the perturbative expansion is valid in the region $`T_KV`$ or $`T_KT`$. We will see that in the steady state, it will be more convenient to introduce a scale $`M_K`$ characterizing both the effects of the thermal energy and of the electric potential driving the current on the Kondo cloud. We will then be able to compare these effects, using a comparison parameter that is exact in one-loop perturbation theory. It is natural that the Callan-Symanzik equation is still valid in the steady state, since the steady state can be understood as an appropriate asymptotic state, characterized by a scale $`V`$, and since averages of operators in asymptotic states satisfy the Callan-Symanzik equation. In particular, $`V`$ should flow trivially with the renormalization group. However, it is instructive to see explicitly how this works in real-time perturbation theory. We will argue that Eq. (83) holds simply from the fact that the Callan-Symanzik equation is satisfied for any average in equilibrium. The main observation is the following. Consider the Hilbert space $`_V`$ associated to the Hamiltonian $`H_0VH_z`$. It is not formed of vectors that are in the Hilbert space $`_0`$ associated to the Hamiltonian $`H_0`$. In particular, its ground state $`|V`$ can be formally defined as $$U_V|0,$$ where $`|0`$ is the ground state of $`H_0`$. This definition indeed makes sense for any finite length of the system, and in fact allows to calculate matrix elements of any operators also at infinite length, but it does not give a vector in $`_0`$ at infinite length. Nevertheless, the mode operators associated to current algebra operators still have a well-defined action on $`_V`$. In order to see this, it is convenient to construct the Hilbert space $`_0`$ by defining a vacuum state $`|0`$ satisfying $`\stackrel{}{J}_{+,,d}(p)|0=0`$ and $`J_z(p)|0=0`$ for all $`p0`$, and by constructing other states of the Hilbert space by acting with $`\stackrel{}{J}_{+,,d}(p)`$ and $`J_z(p)`$ at $`p<0`$. Then, it is a simple matter to see, using (63), that $`\stackrel{}{J}_+(p)|V`$ $`=`$ $`0\text{if and only if}pV`$ $`\stackrel{}{J}_{}(p)|V`$ $`=`$ $`0\text{if and only if}pV`$ $`\stackrel{}{J}_d(p)|V`$ $`=`$ $`0\text{if and only if}p0`$ $`J_z(p)|V`$ $`=`$ $`0\text{if and only if}p>0.`$ (84) The main observation is that any normal ordering operation valid on $`_0`$ is still a good normal ordering operation on $`_V`$. Indeed, a state of the form $$J_+^{i_1}(p_1)\mathrm{}J_+^{i_a}(p_a)J_{}^{j_1}(p_1^{})\mathrm{}J_{}^{j_b}(p_b^{})J_d^{k_1}(p_1^{\prime \prime })\mathrm{}J_d^{k_c}(p_c^{\prime \prime })J_z(p_1^{\prime \prime \prime })\mathrm{}J_z(p_d^{\prime \prime \prime })|V$$ gives zero whenever $$p_1+\mathrm{}+p_a+p_1^{}+\mathrm{}+p_b^{}+p_1^{\prime \prime }+\mathrm{}+p_c^{\prime \prime }+p_1^{\prime \prime \prime }+\mathrm{}+p_d^{\prime \prime \prime }+(ab)V>0.$$ Since $`(ab)V`$ is a much smaller than $`\mathrm{\Lambda }`$ for any finite $`a`$ and $`b`$, a normal ordering of the type $$:J(p_1)J(p_2)\mathrm{}J(p_k):=J(p_{i_1})J(p_{i_2})\mathrm{}J(p_{i_k})\text{with}p_{i_1}p_{i_2}\mathrm{}p_{i_k},$$ (85) where the indices $`i_m`$’s are all different and drawn in appropriate fashion from the set $`\{1,2,\mathrm{},k\}`$ as to make the set of inequalities for the $`p_{i_m}`$’s valid, is still a good normal ordering on $`_V`$. This observation is enough in order to see that the Callan-Symanzik equation is still valid in the steady state. Indeed, the Callan-Symanzik equation is really an equation for operators, rather than just for particular averages. Recalling the regularized real-time perturbation theory (54), consider the operator $`\overline{𝒥}_\mathrm{\Lambda }={\displaystyle \underset{k=0}{\overset{\mathrm{}}{}}}(1)^k{\displaystyle }dp_1R_\mathrm{\Lambda }(p_1)(i\pi \delta (p_1)+\underset{¯}{\mathrm{P}}{\displaystyle \frac{1}{p_1}}){\displaystyle }dp_2R_\mathrm{\Lambda }(p_2)(i\pi \delta (p_1+p_2)+\underset{¯}{\mathrm{P}}{\displaystyle \frac{1}{p_1+p_2}})\times `$ $`\mathrm{}\times {\displaystyle }dp_kR_\mathrm{\Lambda }(p_k)(i\pi \delta (p_1+p_2+\mathrm{}+p_k)+\underset{¯}{\mathrm{P}}{\displaystyle \frac{1}{p_1+p_2+\mathrm{}+p_k}})\times `$ $`{\displaystyle 𝑑sR_\mathrm{\Lambda }(s)[\stackrel{~}{H}_I(p_1),[\stackrel{~}{H}_I(p_2),\mathrm{},[\stackrel{~}{H}_I(p_k),\stackrel{~}{𝒥}(s)]\mathrm{}]]}.`$ (86) It gives the interacting current operator with respect to the free Hilbert space $`_0`$ (at equilibrium) or $`_V`$ (in the steady state). The validity of the Callan-Symanzik equations at equilibrium means that all matrix elements of $`\overline{𝒥}_\mathrm{\Lambda }`$ in the Hilbert space $`_0`$ satisfy the Callan-Symanzik equations. This can be written: $$\left(\mathrm{\Lambda }\frac{}{\mathrm{\Lambda }}|_{\lambda ,\lambda _d}+\beta _\lambda (\lambda ,\lambda _d)\frac{}{\lambda }+\beta _{\lambda _d}(\lambda ,\lambda _d)\frac{}{\lambda _d}\right)\overline{𝒥}_\mathrm{\Lambda }\stackrel{\mathrm{\Lambda }\mathrm{}}{=}0\text{on}_0.$$ (87) Indeed, such matrix elements correspond, in perturbation theory, to matrix elements of the current operator $`𝒥`$ in the basis of eigenstates of the full Hamiltonian $`H|_{V=0}`$ (all of which obey the same Callan-Symanzik equation). We give an argument for the validity of (87) in Appendix C. In (87), the limit $`\mathrm{\Lambda }\mathrm{}`$ should be performed after a matrix element has been calculated, holding the momenta associated to this matrix element fixed. We now extract the divergent and finite part of the operator $`\overline{𝒥}`$ as $`\mathrm{\Lambda }\mathrm{}`$: $$\overline{𝒥}_\mathrm{\Lambda }\underset{j=0}{\overset{\mathrm{}}{}}\frac{1}{j!}\mathrm{ln}^j(\mathrm{\Lambda }):𝒪_j:$$ (88) where $`𝒪_j`$’s are operators built out of the mode operators for the current algebra along with possible impurity-space operators (the normal ordering is defined in (85)). This is always possible to do by rewriting $`\overline{𝒥}_\mathrm{\Lambda }`$ in terms of normal-ordered operators and evaluating the coefficients as $`\mathrm{\Lambda }\mathrm{}`$. The Callan-Symanzik equation (87) then states, $$:𝒪_{j+1}:+(\beta _\lambda (\lambda ,\lambda _d)\frac{}{\lambda }+\beta _{\lambda _d}(\lambda ,\lambda _d)\frac{}{\lambda _d}):𝒪_j:=0(j=0,1,\mathrm{},\mathrm{}).$$ (89) Since the normal ordering operation (85) is also valid on the Hilbert space $`_V`$, it is clear that the expression (88) also gives the divergent and finite part of the current operator on $`_V`$, so that the recursion relation (89) among operators $`:𝒪_j:`$ implies the Callan-Symanzik equation (83) also holds for the steady-state average of the current. Of course, the same is true for any operator that has a well-defined average in the steady state. All this will be explicitly verified for the current operator to one loop in the following sub-section. Density-matrix-independent calculation of the beta functions and of the anomalous dimension of the current to one loop. Let us write $`\overline{𝒥}_\mathrm{\Lambda }`$ using time variables instead of momentum variables: $$\overline{𝒥}_\mathrm{\Lambda }=\underset{k=0}{\overset{\mathrm{}}{}}i^k_{\mathrm{}}^0𝑑t_1_{t_1}^0𝑑t_2\mathrm{}_{t_{k1}}^0𝑑t_k[(H_I^{(0)})_\mathrm{\Lambda }(t_1),[(H_I^{(0)})_\mathrm{\Lambda }(t_2),\mathrm{},[(H_I^{(0)})_\mathrm{\Lambda }(t_k),(𝒥^{(0)})_\mathrm{\Lambda }(0)]\mathrm{}]].$$ (90) Here $$(H_I^{(0)})_\mathrm{\Lambda }(t)=\left(\lambda (\stackrel{}{J}_y)_\mathrm{\Lambda }(t)+\lambda _d(\stackrel{}{J}_d)_\mathrm{\Lambda }(t)\right)\stackrel{}{S}$$ (91) and $$(𝒥^{(0)})_\mathrm{\Lambda }(t)=\lambda (\stackrel{}{J}_x)_\mathrm{\Lambda }(t)\stackrel{}{S}.$$ (92) In the regularization scheme that we consider, characterized by the regulator $`R_\mathrm{\Lambda }(p)`$ (50), it is a simple matter to observe that $$\mathrm{\Lambda }\frac{}{\mathrm{\Lambda }}(H_I^{(0)})_\mathrm{\Lambda }(t)=\frac{1}{\mathrm{\Lambda }^2}(H_I^{(0)})_\mathrm{\Lambda }^{\prime \prime }(t),\mathrm{\Lambda }\frac{}{\mathrm{\Lambda }}(𝒥^{(0)})_\mathrm{\Lambda }(t)=\frac{1}{\mathrm{\Lambda }^2}(𝒥^{(0)})_\mathrm{\Lambda }^{\prime \prime }(t)$$ (93) where primes mean time derivatives. Consider the first few terms of (90): $`\overline{𝒥}_\mathrm{\Lambda }`$ $`=`$ $`(𝒥^{(0)})_\mathrm{\Lambda }+{\displaystyle _{\mathrm{}}^0}𝑑t[i(H_I^{(0)})_\mathrm{\Lambda }(t),(𝒥^{(0)})_\mathrm{\Lambda }(0)]+`$ (94) $`+{\displaystyle _{\mathrm{}}^0}𝑑t_1{\displaystyle _{t_1}^0}𝑑t_2[i(H_I^{(0)})_\mathrm{\Lambda }(t_1),[i(H_I^{(0)})_\mathrm{\Lambda }(t_2),(𝒥^{(0)})_\mathrm{\Lambda }(0)]]+\mathrm{}`$ Using integration by parts, we find $`\mathrm{\Lambda }{\displaystyle \frac{}{\mathrm{\Lambda }}}\overline{𝒥}`$ $`=`$ $`{\displaystyle \frac{1}{\mathrm{\Lambda }^2}}((𝒥^{(0)})_\mathrm{\Lambda }^{\prime \prime }(0)+`$ (95) $`+[i(H_I^{(0)})_\mathrm{\Lambda }^{}(0),(𝒥^{(0)})_\mathrm{\Lambda }(0)]+{\displaystyle _{\mathrm{}}^0}𝑑t[i(H_I^{(0)})_\mathrm{\Lambda }(t),(𝒥^{(0)})_\mathrm{\Lambda }^{\prime \prime }(0)]+`$ $`+{\displaystyle _{\mathrm{}}^0}dt[[i(H_I^{(0)})_\mathrm{\Lambda }^{}(t),i(H_I^{(0)})_\mathrm{\Lambda })(t)],(𝒥^{(0)})_\mathrm{\Lambda }(0)]+`$ $`+{\displaystyle _{\mathrm{}}^0}𝑑t[(H_I^{(0)})_\mathrm{\Lambda }(t),[(H_I^{(0)})_\mathrm{\Lambda }^{}(0),(𝒥^{(0)})_\mathrm{\Lambda }(0)]]+`$ $`+{\displaystyle _{\mathrm{}}^0}dt_1{\displaystyle _{t_1}^0}dt_2[i(H_I^{(0)})_\mathrm{\Lambda }(t_1),[i(H_I^{(0)})_\mathrm{\Lambda }(t_2),(𝒥^{(0)})_\mathrm{\Lambda }^{\prime \prime }(0)]]+\mathrm{}).`$ We want to evaluate all this at $`\mathrm{\Lambda }\mathrm{}`$ in order to find the beta function using the equation (87). Inside the parenthesis, we need only keep the contributions of order $`\mathrm{\Lambda }^2\mathrm{ln}^j(\mathrm{\Lambda })`$ for non-negative integers $`j`$. The very first operator of course does not contribute, but all others do. These leading contributions have to be compared with the leading contributions of $$\left(\beta _\lambda \frac{}{\lambda }+\beta _{\lambda _d}\frac{}{\lambda _d}\right)\overline{𝒥}=\left(\beta _\lambda \frac{}{\lambda }+\beta _{\lambda _d}\frac{}{\lambda _d}\right)\left[(𝒥^{(0)})_\mathrm{\Lambda }(0)+_{\mathrm{}}^0𝑑t[i(H_I^{(0)})_\mathrm{\Lambda }(t),(𝒥^{(0)})_\mathrm{\Lambda }(0)]+\mathrm{}\right]$$ (96) where we wrote only the terms contributing to the one-loop order. The beta functions appearing there can be obtained by requiring that when the derivatives with respect to the couplings are applied to the operator $`i(H_I^{(0)})_\mathrm{\Lambda }(t)`$, they give the operator $`[i(H_I^{(0)})_\mathrm{\Lambda }^{}(t),i(H_I^{(0)})_\mathrm{\Lambda })(t)]`$ (appearing on the third line of (95)) in the limit $`\mathrm{\Lambda }\mathrm{}`$. The main contribution in this limit of this operator can be obtained from $$\frac{1}{\mathrm{\Lambda }^2}[i(H_I^{(0)})_\mathrm{\Lambda }^{}(t),i(H_I^{(0)})_\mathrm{\Lambda })(t)]=i((\lambda _d^2+\lambda ^2)(\stackrel{}{J}_d)_{\sqrt{2}\mathrm{\Lambda }}(t)+2\lambda \lambda _d(\stackrel{}{J}_y)_{\sqrt{2}\mathrm{\Lambda }}(t))\stackrel{}{S}+\frac{1}{\mathrm{\Lambda }^2}:𝒪:$$ (97) where $`𝒪`$ contains products of current algebra operators at the same point $`x=t`$; the explicit form of $`𝒪`$ is not important here. Equating the leading behavior of this operator as $`\mathrm{\Lambda }\mathrm{}`$ with that of $$\left(\beta _\lambda \frac{}{\lambda }+\beta _{\lambda _d}\frac{}{\lambda _d}\right)iH_I^{(0)}(t)$$ gives $$\beta _{\lambda _d}=(\lambda _d^2+\lambda ^2)+O(\lambda _d^3,\lambda ^4,\lambda _d^2\lambda ^2),\beta _\lambda =2\lambda \lambda _d+O(\lambda \lambda _d^2,\lambda ^3).$$ (98) Note that we had to take the limit $`\mathrm{\Lambda }\mathrm{}`$ of the operator $`[i(H_I^{(0)})_\mathrm{\Lambda }^{}(t),i(H_I^{(0)})_\mathrm{\Lambda })(t)]`$ rather than of the integral where it is involved on the third line of (95), since the beta function does not depend on the particular average that we are calculating. The sub-leading operators in (97) may give contributions to this integral, but these are two-loop contributions to the anomalous dimension of the operator $`𝒥`$ (of course, since this operator should have zero anomalous dimensions, all such contributions should cancel out). The total one-loop contributions to the anomalous dimension of $`𝒥`$ can be verified to be zero by checking that when the derivatives with respect to the couplings in (96) are applied to the operator $`(𝒥^{(0)})_\mathrm{\Lambda }(0)`$ (the first term inside the parenthesis), they give the two terms appearing on the second line of (95) in the limit $`\mathrm{\Lambda }\mathrm{}`$. In this limit, the second term on the second line of (95) can be written $$_{\mathrm{}}^0𝑑t[i(H_I^{(0)})_\mathrm{\Lambda }(t),𝒥^{\prime \prime }]_{\mathrm{}}^0𝑑t[i(H_I^{(0)})_\mathrm{\Lambda }^{\prime \prime }(t),𝒥]=[i(H_I^{(0)})_\mathrm{\Lambda }^{}(0),𝒥].$$ (99) That is, it is equal to the first term. Together, their leading behavior at $`\mathrm{\Lambda }\mathrm{}`$ can be obtained from $$\frac{2}{\mathrm{\Lambda }^2}[i(H_I^{(0)})_\mathrm{\Lambda }^{}(0),(𝒥^{(0)})_\mathrm{\Lambda }(0)]=2\lambda _d(𝒥^{(0)})_{\sqrt{2}\mathrm{\Lambda }}(0)+\frac{1}{\mathrm{\Lambda }^2}:\stackrel{~}{𝒪}:$$ (100) where $`\stackrel{~}{𝒪}`$ is of the same form as $`𝒪`$. But since $$\left(\beta _\lambda \frac{}{\lambda }+\beta _{\lambda _d}\frac{}{\lambda _d}\right)𝒥=\lambda ^1\beta _\lambda 𝒥=2\lambda _d𝒥+O(\lambda \lambda _d^2,\lambda ^3),$$ we immediately conclude that the anomalous dimension of the current is zero to one loop. It is a simple matter to verify that the one-loop steady-state current (68) with the values (69) and (73) indeed satisfies the Callan-Symanzik equation $$\left(\mathrm{\Lambda }\frac{}{\mathrm{\Lambda }}|_{\lambda ,\lambda _d}2\lambda \lambda _d\frac{}{\lambda }(\lambda _d^2+\lambda ^2)\frac{}{\lambda _d}\right)𝒥_{ss}\stackrel{\mathrm{\Lambda }\mathrm{}}{=}0.$$ (101) Note that our derivation of the Callan-Symanzik equation to one-loop did not use the particular initial density matrix $`e^{\beta (H_0VH_z)}`$; only the fact that normal-ordered operators are finite when averaged with this density matrix. Two-loop beta functions. The two-loop result (80) with two-loop divergent parts (81) and one-loop finite and divergent parts (69) and (73) is $`𝒥_{ss}`$ $``$ $`3\pi V\lambda ^2({\displaystyle \frac{1}{2}}+2\lambda _d(P(w)+\mathrm{ln}\left({\displaystyle \frac{\mathrm{\Lambda }}{V}}\right))+`$ (102) $`\lambda _d^2[10P(w)+5\mathrm{ln}\left({\displaystyle \frac{\mathrm{\Lambda }}{V}}\right)1]\mathrm{ln}\left({\displaystyle \frac{\mathrm{\Lambda }}{V}}\right)+\lambda ^2[2P(w)+\mathrm{ln}\left({\displaystyle \frac{\mathrm{\Lambda }}{V}}\right)1]\mathrm{ln}\left({\displaystyle \frac{\mathrm{\Lambda }}{V}}\right)).`$ This satisfies perturbatively the Callan-Symanzik equation with beta functions $`\beta _{\lambda _d}`$ $`=`$ $`\lambda _d^2\lambda ^2+O(\lambda _d^3,\lambda ^2\lambda _d)`$ $`\beta _\lambda `$ $`=`$ $`2\lambda \lambda _d+\lambda \lambda _d^2+\lambda ^3+O(\lambda \lambda _d^3,\lambda ^3\lambda _d).`$ (103) Note that this calculation does not give the third order coefficient of $`\beta _{\lambda _d}`$. However, we know the universal beta functions of the one-channel Kondo model $`(g^2\frac{1}{2}g^3)`$ and that of the symmetric two-channel Kondo model $`(g^2g^3)`$. Taking $`V=0`$ (which does not affect the beta function), these two cases are obtained respectively at $`\lambda =0`$, and at $`\lambda _d=\lambda `$ upon diagonalization of the matrix of couplings $`J_{\alpha ,\alpha ^{}}`$. These two facts essentially fix the two-loop beta function to be of the form $`\beta _{\lambda _d}`$ $`=`$ $`(\lambda _d^2+\lambda ^2\lambda _d^3\lambda ^2\lambda _d+\mathrm{})`$ $`\beta _\lambda `$ $`=`$ $`(2\lambda \lambda _da\lambda \lambda _d^2(2a)\lambda ^3+\mathrm{})`$ (104) where $`a`$ is a non-universal number. Our results fix $`a=1`$, which gives the standard beta functions $`\beta _{\lambda _d}`$ $`=`$ $`(\lambda _d^2+\lambda ^2\lambda _d^3\lambda ^2\lambda _d+O(\lambda _d^4,\lambda ^2\lambda _d^2,\lambda ^4))`$ $`\beta _\lambda `$ $`=`$ $`(2\lambda \lambda _d\lambda \lambda _d^2\lambda ^3+O(\lambda \lambda _d^3,\lambda ^3\lambda _d)).`$ (105) It is convenient now to change variables to $`\lambda _\pm =\lambda _d\pm \lambda `$ so that (to the same order) $$\beta _\pm =\lambda _\pm ^2+\frac{1}{2}\lambda _\pm (\lambda _+^2+\lambda _{}^2).$$ (106) The RG invariant anisotropy parameter can be expressed as: $$C=\frac{1}{2}\left(\frac{1}{\lambda _{}}\left(1\frac{1}{2}\lambda _+\right)\frac{1}{\lambda _+}\left(1\frac{1}{2}\lambda _{}\right)\right)$$ (107) The scaling limit of the current. We now evaluate the current as function of the voltage and of the temperature. For the expression of the current, our analysis will make use solely of the one-loop results (77). Expression (77) is the perturbative expansion in the limit $`T,V\mathrm{\Lambda }`$ of the model. The couplings $`\lambda _d,\lambda `$ take their “microscopic” values: the actual values for the physical process represented by the interaction terms, and the cut-off $`\mathrm{\Lambda }`$ is of the scale of the band width. Of course, since we neglected terms vanishing as $`\mathrm{\Lambda }\mathrm{}`$, this is not an exact expression for finite $`\mathrm{\Lambda }`$. However, it allows us to have an exact expression in the scaling limit. As explained earlier, we expect universality for $`T,V\mathrm{\Lambda }`$. It is convenient therefore to consider the limit where $`\mathrm{\Lambda }`$ is sent to infinity with the couplings becoming cut-off dependent: they are modified so as to keep the physics unchanged when the cut-off is increased. Denoting for the moment $`\mathrm{\Lambda }=\mathrm{\Lambda }_0`$ the physical value of the cut-off (of the order of the band width) and $`\lambda =\lambda ^0,\lambda _d=\lambda _d^0`$ the values of the coupling there, the running couplings $`\lambda ^r(\mathrm{\Lambda }),\lambda _d^r(\mathrm{\Lambda })`$, are governed by the RG equations, $`\mathrm{\Lambda }{\displaystyle \frac{d\lambda _d^r}{d\mathrm{\Lambda }}}`$ $`=`$ $`(\lambda _d^r)^2+(\lambda ^r)^2(\lambda _d^r)^3(\lambda ^r)^2\lambda _d^r`$ $`\mathrm{\Lambda }{\displaystyle \frac{d\lambda ^r}{d\mathrm{\Lambda }}}`$ $`=`$ $`2\lambda ^r\lambda _d^r\lambda ^r(\lambda _d^r)^2(\lambda ^r)^3`$ (108) with initial conditions fixed by the microscopic values of the couplings: $`\lambda _d^r(\mathrm{\Lambda }=\mathrm{\Lambda }_0)=\lambda _d^0,\lambda ^r(\mathrm{\Lambda }=\mathrm{\Lambda }_0)=\lambda ^0`$. The solution of the RG flow can then be described by RG invariants \- quantities that describe the full trajectory. Such quantities are $`C`$ introduced earlier, and a scale $`T_K`$ to be discussed below. Thus any physical quantity $`F`$ will depend on the cut off and coupling via these invariants $`F=F(T/T_K,V/T_K,C)`$. One may reformulate this scaling procedure as follows. Introduce the scale $`M`$, the physical scale on which the system is examined, $$V=M\mathrm{sin}(\alpha ),T=M\mathrm{cos}(\alpha )$$ for some angle $`\alpha `$ in the $`VT`$ plane so that the previous result for the current is written as, $$𝒥_{ss}=\frac{3\pi }{2}V\lambda ^2\left[1+4\lambda _d\left(P(\mathrm{tan}(\alpha ))+\mathrm{ln}\left(\frac{\mathrm{\Lambda }}{M}\mathrm{csc}(\alpha )\right)\right)+\mathrm{}\right]$$ Following the previous considerations, the current can be written in terms of couplings that depend on the physical scale $`M`$, satisfying the equations with respect to $`M`$, $`M{\displaystyle \frac{d\lambda _d^r}{dM}}`$ $`=`$ $`(\lambda _d^r)^2(\lambda ^r)^2+(\lambda _d^r)^3+(\lambda ^r)^2\lambda _d^r`$ $`M{\displaystyle \frac{d\lambda ^r}{dM}}`$ $`=`$ $`2\lambda ^r\lambda _d^r+\lambda ^r(\lambda _d^r)^2+(\lambda ^r)^3`$ (109) with initial conditions $`\lambda _d^r(M=\mathrm{\Lambda })=\lambda _d,\lambda ^r(M=\mathrm{\Lambda })=\lambda `$ ($`\mathrm{\Lambda }`$ being of the order of the band width), as follows: $$𝒥_{ss}=\frac{3\pi }{2}V(\lambda ^r)^2\left[1+4\lambda _d^rQ(\alpha )+\mathrm{}\right]$$ (110) where $$Q(\alpha )=P(\mathrm{tan}(\alpha ))+\mathrm{ln}(\mathrm{csc}(\alpha )).$$ (111) Again, the solution to the RG flow (IV) should be described solely as a function of $`M/T_K`$ and of the RG invariant $`C`$ (107), instead of $`\mathrm{\Lambda }`$ and the initial conditions $`\lambda _d`$ and $`\lambda `$. With such a description, we can trivially take the scaling limit: $`T,V,T_K\mathrm{\Lambda }`$ with fixed ratios $`T:V:T_K`$ and fixed $`C`$, since $`\mathrm{\Lambda }`$ does not appear anymore. In this limit, the quantum field theory gives exact results, and the system is in its universal regime. In order to have unambiguous results, we need to define $`T_K`$. Once $`T_K`$ is defined, one need only solve the RG flow (IV) (a numerical solution is easy to obtain with good precision, for instance), and one obtains the scaling limit in its perturbative region. The actual values of $`\lambda _d,\lambda `$ and of the band width $`\mathrm{\Lambda }`$ should be such that the theory is near to the scaling limit if we want the system to be meaningfully described by quantum field theory. The perturbative region of the scaling limit is $`T_K\sqrt{V^2+T^2}`$, for any value of $`C`$. Hence, in order to define $`T_K`$ in perturbation theory, we need to look at the expansion as $`M/T_K\mathrm{}`$ of the solution to the RG flow (IV). This expansion, and the value of $`T_K`$, have different forms depending on $`C`$. We will consider here two cases: taking $`C=\mathrm{}`$ then $`M/T_K\mathrm{}`$; and keeping $`C`$ finite then taking $`M/T_K\mathrm{}`$. The first case corresponds to $`\lambda _d=\lambda `$. In fact, it is true to all orders that if $`\lambda _d=\lambda `$, then $`\lambda _d^r=\lambda ^r`$ for all scales $`M`$. In this case, the equilibrium Hamiltonian $`H|_{V=0}`$ is in fact a one-channel Kondo model plus a decoupled free massless fermion. The case $`C`$ finite corresponds naturally to $`\lambda C\lambda _d^2`$ in the scaling limit (as will become clear below). In the case where $`\lambda _d=\lambda `$, the RG equations lead to the following large-($`M/T_K`$) expansion of the running couplings: $$\lambda _d^r=\lambda ^r=\frac{1}{2\mathrm{ln}(M/T_K)}+\frac{\mathrm{ln}\mathrm{ln}(M/T_K)}{4\mathrm{ln}^2(M/T_K)}+\frac{a}{2\mathrm{ln}^2(M/T_K)}+\frac{\mathrm{ln}^2\mathrm{ln}(M/T_K)}{8\mathrm{ln}^3(M/T_K)}+\frac{(4a1)\mathrm{ln}\mathrm{ln}(M/T_K)}{8\mathrm{ln}^3(M/T_K)}+O\left(\frac{1}{\mathrm{ln}^3(M/T_K)}\right).$$ (112) The coefficient $`a`$ can be changed by a change of the scale (the integration constant) $`T_K`$; making the replacement $`T_KxT_K`$ is equivalent to doing $`aa+\mathrm{ln}(x)`$. Note also that a change of scale $`T_KxT_K`$ corresponds simply to a perturbative change of the running coupling constants that keeps the beta functions invariant. Fixing the value of $`a`$ is making a choice of definition for $`T_K`$, which is one more condition necessary to totally specify the renormalization procedure. Of course, different definitions of $`T_K`$ reproduce the same scaling limit. For arbitrary $`a`$, in terms of the coupling $`\lambda _d=\lambda `$ and of the scale of the band width $`\mathrm{\Lambda }`$, $`T_K`$ has the form $$T_K(a)=\mathrm{\Lambda }\sqrt{2\lambda }e^{a\frac{1}{2\lambda }}(1+O(\lambda ))(\lambda _d=\lambda ).$$ A standard definition for the Kondo temperature is to make the term in $`1/\mathrm{ln}^2(M/T_K)`$ vanish ($`a=0`$), giving: $$T_K=\mathrm{\Lambda }\sqrt{2\lambda }e^{\frac{1}{2\lambda }}(1+O(\lambda ))(\lambda _d=\lambda ).$$ (113) In the case where the RG invariant $`C`$ is finite, the large-($`M/T_K`$) expansion has the form $`\lambda _d^r`$ $`=`$ $`{\displaystyle \frac{1}{\mathrm{ln}(M/T_K)}}+{\displaystyle \frac{\mathrm{ln}\mathrm{ln}(M/T_K)}{\mathrm{ln}^2(M/T_K)}}+{\displaystyle \frac{a}{\mathrm{ln}^2(M/T_K)}}+{\displaystyle \frac{\mathrm{ln}^2\mathrm{ln}(M/T_K)}{\mathrm{ln}^3(M/T_K)}}+{\displaystyle \frac{(2a1)\mathrm{ln}\mathrm{ln}(M/T_K)}{\mathrm{ln}^3(M/T_K)}}+\mathrm{}`$ $`{\displaystyle \frac{\lambda ^r}{C}}`$ $`=`$ $`{\displaystyle \frac{1}{\mathrm{ln}^2(M/T_K)}}+{\displaystyle \frac{2\mathrm{ln}\mathrm{ln}(M/T_K)}{\mathrm{ln}^3(M/T_K)}}+{\displaystyle \frac{1+2a}{\mathrm{ln}^3(M/T_K)}}+{\displaystyle \frac{3\mathrm{ln}^2\mathrm{ln}(M/T_K)}{\mathrm{ln}^4(M/T_K)}}+{\displaystyle \frac{(6a+1)\mathrm{ln}\mathrm{ln}(M/T_K)}{\mathrm{ln}^4(M/T_K)}}+\mathrm{}`$ where the dots ($`\mathrm{}`$) mean $`O(\mathrm{ln}^3(M/T_K))`$ for $`\lambda _d^r`$, and $`O(\mathrm{ln}^4(M/T_K))`$ for $`\lambda ^r`$. Note that in general, $`C`$ will not appear only as a normalization of $`\lambda `$; this is an artifact of the limited perturbative order which we consider. Also, note that $`C`$ must be positive for $`\lambda ^r`$ to be positive. Again, a variation of $`T_K`$ has the effect of changing $`a`$: making the replacement $`T_KxT_K`$ is equivalent to doing $`aa+\mathrm{ln}(x)`$. For arbitrary $`a`$, we can use the first or the second equation of (LABEL:largeM) in order to determine $`T_K`$ in terms of the couplings $`\lambda _d,\lambda `$ and of the scale of the band width $`\mathrm{\Lambda }`$. This gives two equivalent forms: $$T_K(a)=\mathrm{\Lambda }\lambda _de^{a\frac{1}{\lambda _d}}(1+O(\lambda _d))=\mathrm{\Lambda }\sqrt{\frac{\lambda }{C}}e^{a\frac{1}{2}\sqrt{\frac{C}{\lambda }}}(1+O(\sqrt{\lambda }))(\lambda _d\lambda ).$$ Recall that from (107), we indeed have that $`\lambda \lambda _d^2`$ in the scaling limit, so that the second equality above is correct. A more standard way of writing the Kondo temperature can be obtained by considering the linear combination $`\lambda _d^r+\lambda ^r`$. This gives, again for the same object $`T_K(a)`$, $$T_K(a)=\mathrm{\Lambda }(\lambda _d+\lambda )e^{(a+C)\frac{1}{\lambda _d+\lambda }}(1+O(\lambda _d))(\lambda _d\lambda ).$$ A standard definition is $`a=C`$, giving $$T_K=\mathrm{\Lambda }(\lambda _d+\lambda )e^{\frac{1}{\lambda _d+\lambda }}(1+O(\lambda _d))(\lambda _d\lambda ).$$ (115) Four comments are now in order. First, note that taking $`C\mathrm{}`$ in (LABEL:largeM) does not give (112). This is expected, since the former corresponds to taking first $`M/T_K\mathrm{}`$ then $`C\mathrm{}`$ in the solution to the RG equations, whereas the latter corresponds to taking first $`C\mathrm{}`$ then $`M/T_K\mathrm{}`$, and these two limits do not commute. In particular, one would have obtained still different expressions taking simultaneously $`C\mathrm{}`$ and $`M/T_K\mathrm{}`$, with a prescribed ratio between them, for instance (it is a simple matter to evaluate in this case, or in any other case, the expansion as $`M/T_K\mathrm{}`$ of the solutions to the RG equations). The expressions (LABEL:largeM) and (112) should not be regarded as describing all solutions to the RG equation; they are rather particular limits of the solutions, as described above, and are used here solely for the purpose of defining the Kondo temperature purely using perturbation theory (the perturbative regime is, again, the regime $`MT_K`$, for any $`C`$ taken to infinity or not, simultaneously or not with $`M/T_K`$). The expansion (LABEL:largeM) offers a way of defining $`T_K`$ in the regime with $`C`$ finite, whereas the expansion (112) offers a way of defining it in the regime with $`C=\mathrm{}`$. Other definitions would have been possible, but we only need these two definitions of the Kondo temperature here. Second, it is important to recall that the cases $`C=\mathrm{}`$ and $`C<\mathrm{}`$ do exhaust all possible scaling regimes, even though the expressions (112) and (LABEL:largeM) do not exhaust all possible behavior of the running couplings as $`M/T_K\mathrm{}`$. However, there are many ways of reaching any given scaling regime. In particular, the relations $`\lambda =\lambda _d`$ and $`\lambda C\lambda _d^2`$ (naturally associated to, respectively, the scale definitions (113) and (115)) do not exhaust all possible ways the same scaling limit can be reached. In order to understand what this means, first recall that the full RG-improved perturbation theory reproduces the divergent and finite part of the full bare perturbation theory, so that we can talk about the RG trajectory for the bare couplings before taking the scaling limit. Then, from this viewpoint, one can take the scaling limit by fixing a trajectory $`C`$, and by sending $`\mathrm{\Lambda }\mathrm{}`$ while keeping $`\lambda _d`$ and $`\lambda `$ on the trajectory at scale $`\mathrm{\Lambda }`$. For finite $`C`$, this gives $`\lambda C\lambda _d^2`$, whereas for $`C=\mathrm{}`$, this gives $`\lambda =\lambda _d`$. But one could also take the scaling limit by changing the value of $`C`$ (changing the shape of the trajectory) while $`\mathrm{\Lambda }\mathrm{}`$, always keeping $`\lambda _d`$ and $`\lambda `$ on the trajectory defined by $`C`$. For instance, this is what happens if one takes $`\lambda =q\lambda _d`$ for some fixed $`q1`$ when sending $`\lambda _d,\lambda 0`$; then one must simultaneously take $`C\mathrm{}`$ in order to keep $`\lambda `$ and $`\lambda _d`$ on the RG trajectory. The resulting value of $`T_K`$ in terms of such couplings is different from (113) and (115). But in this example, when the scaling limit is reached we have $`C=\mathrm{}`$, so that the quantum field theory is the same as the one obtained by taking $`\lambda =\lambda _d`$ and sending them to 0; in particular, in the scaling limit, we still have $`\lambda ^r=\lambda _d^r`$, and we can still define a scale $`T_K`$ using (112). Since we are only interested in the scaling regimes, we do not need to look at all possible ways a given regime can be reached. Third, it is important to note that the leading behavior of the current at large $`M/T_K`$ is very different if $`C=\mathrm{}`$ or if $`C<\mathrm{}`$ (again, we only look at the two cases mentionned above). In the first case, it is given by $$𝒥_{ss}\frac{3\pi }{8}\frac{M\mathrm{sin}(\alpha )}{\mathrm{ln}^2(M/T_K)}=\frac{3\pi }{8}\frac{V}{\mathrm{ln}^2(\sqrt{V^2+T^2}/T_K)}(\lambda _d=\lambda ).$$ (116) On the other hand, in the second case it is given by $$𝒥_{ss}\frac{3\pi C^2}{2}\frac{M\mathrm{sin}(\alpha )}{\mathrm{ln}^4(M/T_K)}=\frac{3\pi C^2}{2}\frac{V}{\mathrm{ln}^4(\sqrt{V^2+T^2}/T_K)}(\lambda _d\lambda ).$$ (117) Note that since the RG invariant $`C`$ appears as a coefficient, the leading behavior in this case is non-universal; this is a property of the regime $`C`$ finite. Finally, the parameter $`\alpha `$ can be seen as parametrizing a family of choices of infrared cutoffs for our theory. For instance, at $`\alpha =0`$, the temperature is the infrared cutoff, whereas at $`\alpha =\pi /2`$, the voltage solely plays the role of an infrared cutoff. It is important to note that the voltage is a good infrared cutoff for the average of the current operator to two-loop order. Indeed, the limit $`T/V0`$ of our one-loop and two-loop bare perturbative results is finite; equivalently, the limit $`\alpha \pi /2`$ of the renormalized perturbative results is finite. We should remark however that this need not be the case when other quanties are considered, see rg1 . Universal ratios. From the viewpoint of the interpretation of the measurement of non-equilibrium quantities, the Kondo temperature, as defined for instance in (113), is not very convenient, since the temperature $`T`$ and the bias voltage $`V`$ have a different influence on the Kondo screening cloud, and we would like a definition that embodies this difference. Hence, it seems appropriate to define a continuum of scales, $`M_K(\alpha )`$, depending on the angle $`\alpha `$ on the $`VT`$ plane. Consider first the case $`\lambda _d=\lambda `$. A possible definition of $`M_K(\alpha )`$ is the requirement that the average current $`𝒥`$ in the steady state does not have a term of the form $`M/\mathrm{ln}^3(M/M_K(\alpha ))`$ in its expansion at large $`M/M_K(\alpha )`$: $$𝒥=\frac{3\pi }{8}\frac{M\mathrm{sin}(\alpha )}{\mathrm{ln}^2(M/M_K(\alpha ))}\left(1+\frac{\mathrm{ln}\mathrm{ln}(M/M_K(\alpha ))}{\mathrm{ln}(M/M_K(\alpha ))}+O\left(\frac{\mathrm{ln}^2\mathrm{ln}(M/M_K(\alpha ))}{\mathrm{ln}^4(M/M_K(\alpha ))}\right)\right).$$ (118) This gives: $$M_K(\alpha )=T_Ke^{Q(\alpha )}$$ (119) where $`T_K`$ is defined in (113). As particular cases, we can now define the “decoherence” Kondo scale, $`M_K(\pi /2)`$, as the voltage at $`T=0`$ at which the Kondo cloud is destroyed by the electrons passing through, and the “thermal” Kondo scale $`M_K(0)`$. We find that their ratio, which is universal and does not receive corrections from the two-loop (or higher-loop) contributions to the beta function (nor from higher order contributions to the current), is $$R_{\text{current}}=\frac{M_K(\pi /2)}{M_K(0)}=2\pi e^{1\frac{3\gamma }{2}\frac{\sqrt{\pi }}{2}}=2.96188723\mathrm{}$$ (120) In the case $`\lambda _d\lambda `$ (that is, for $`C<\mathrm{}`$), a similar definition of $`M_K(\alpha )`$ can be made: requiring that the current does not have a term of the form $`M/\mathrm{ln}^4(M/M_K(\alpha ))`$. Constructing again the ratio $`M_K(\pi /2)/M_K(0)`$ gives exactly the same number: this ratio is indeed universal, independent of both $`T_K`$ and $`C`$. It is important that the current possess an infrared-convergent (convergence at large switch-on times) perturbative expansion to one loop for these universal ratios to have a meaning. Further, that they have the usual one-loop logarithmic accuracy near to the scaling limit is a consequence of the infrared convergence of the two-loop perturbation theory. The scale $`M_K(\alpha )`$ defined above is characteristic of the current; other physical quantities would give different functions $`M_K(\alpha )`$, and different ratios. For instance, the same analysis can be applied on the differential conductance $$G=\frac{d}{dV}𝒥_{ss}.$$ From the perturbative calculations, we have $$G=\frac{3\pi }{2}\lambda ^2\left[1+4\lambda _d\left(P\left(\frac{V}{T}\right)+\mathrm{ln}\left(\frac{\mathrm{\Lambda }}{V}\right)+\frac{V}{T}P^{}\left(\frac{V}{T}\right)1\right)+\mathrm{}\right].$$ (121) Again, in terms of running couplings, we have $$G=\frac{3\pi }{2}(\lambda ^r)^2\left[1+4\lambda _d^r\stackrel{~}{Q}(\alpha )+\mathrm{}\right]$$ (122) where $$\stackrel{~}{Q}(\alpha )=P(\mathrm{tan}(\alpha ))+\mathrm{ln}(\mathrm{csc}(\alpha ))+\mathrm{tan}(\alpha )P^{}(\mathrm{tan}(\alpha ))1.$$ (123) Taking again $`\lambda _d=\lambda `$, we can repeat the calculations above and define similarly the scale $`\stackrel{~}{M}_K(\alpha )`$ associated to the conductance. We find $$\stackrel{~}{M}_K(\alpha )=T_Ke^{\stackrel{~}{Q}(\alpha )}.$$ (124) We observe that $`\stackrel{~}{Q}(0)=Q(0)`$ and that $`\stackrel{~}{Q}(\pi /2)=Q(\pi /2)1`$. Hence, we have $$R_{\text{conductance}}=\frac{\stackrel{~}{M}_K(\pi /2)}{\stackrel{~}{M}_K(0)}=e^1R_{\text{current}}=2\pi e^{\frac{3\gamma }{2}\frac{\sqrt{\pi }}{2}}=1.08961742\mathrm{}$$ (125) ## V Perspectives Reaching the steady-state. We showed that the bath of free massless fermions suffices to allow the system to reach equilibrium at temperature $`T`$ in the case of zero bias voltage, and to allow it to reach steady state when the bias voltage is non-zero. No other relaxation process has to be assumed; the infinite bath of free massless fermions plays the role of a thermal bath and is able to absorb the energy necessary for relaxation to occur, as well as for the steady state to exist. This is in close connection with the study of Caldeira and Leggett CaldeiraL81 : they constructed a model where an infinite number of oscillators provides an explicit dissipation in order to study, from first principles, the effect of dissipation on quantum tunneling. It turns out CallanT90 that their construction is simply related to models of field theory that are conformal in the bulk (in general, with non-conformal boundary conditions), as recall in the Introduction. It would be interesting to see to what extent our proof can be generalized to the study of more general models of quantum kinetics with thermal dissipation, and to see whether similar arguments can be applied to other impurity models out of equilibrium. A principle to follow, as can be extracted from our derivation, is the fact that the orbit of a global symmetry of the model should cover the possible values of the boundary degrees of freedom. Related to the latter point, another question is: What is the effect of a magnetic field on the steady state in the quantum dot? Our proof of factorization in Appendix B does not hold anymore when a magnetic field is present, since it uses heavily the invariance of the correlation functions under $`SU(2)`$ transformations. Hence, the real-time perturbation series is no longer expected to be convergent as the switch-on time is sent to minus infinity. This is simple to understand physically: at small couplings, both the current through the dot and its interaction with the leads are weak, that is, both the non-equilibrating effect and the thermalization effect are weak. At zero magnetic field, since all states of the isolated dot have the same energy, the thermalization effect is more efficient and still stronger. At non-zero magnetic field, however, as the couplings are sent to zero, we cannot expect that the dot smoothly reaches its thermal equilibrium energy distribution when no other thermal bath is coupled to it. The real-time perturbation theory should describe this situation, but obviously its zeroth order cannot give anything else than the thermal equilibrium value of any quantity under study. Hence, in non-zero magnetic field, we can expect some strong non-analyticity in the couplings and we must find large-time (IR) divergences in the perturbative coefficients. This clearly indicates that the real-time perturbative series does not properly describe the approach to the steady state, neither the steady state itself, of the model in magnetic field without external thermal bath. Questions remain, as raised in the Introduction: Does the quantum field theory still (non-perturbatively) reach a steady state, or does it show other behaviors at large times, like oscillations (of the dot magnetization, for instance)? If it reaches a steady state, is it a good description of realistic systems, where the dot is coupled to an external thermal bath at all times (without exchange of particles), independently of its coupling to the leads? The former question was partially answered, in rg6 : it was assumed on physical grounds that the non-equilibrium Kondo model reaches a steady state, and mainly from this assumption, it was explained how to obtain the zeroth order of perturbation theory for the dot magnetization. Indeed, strong non-analyticity is obtained. In a sense, one should start with a density matrix that already contains a non-thermal distribution of the impurity spin states; this density matrix can be obtained by requiring that the perturbative series be convergent at large times. As explained clearly there, this is equivalent to solving a quantum Boltzmann equation in order to determine the non-thermal dot occupation numbers. This should answer partly, in some sense still perturbatively, the question of describing the steady state of the quantum field theory (the non-equilibrium Kondo model with magnetic field) – although the results really start with the assumption of a steady state, and do not establish its existence. However, this does not address the question as to whether the leads correctly play the role of thermal baths, or whether a coupling to an external thermal bath would have important effects. In relation to the latter question, our argument suggests that, at least at small couplings (or at temperatures much greater than the Kondo temperature), the model does not describe the true steady state of the non-equilibrium Kondo dot in contact with a thermal environment. Indeed, from a physical interpretation of the perturbative series, the leads do not provide a strong enough thermalisation to absorb the energy necessary for the steady state to occur, and it is possible that it cannot be trusted to sustain the correct steady state. That it does not provide the thermalisation for the steady state to occur is certainly in agreement with rg6 : there it was one of the main points that the large-time divergences are due to the absence of a proper thermal bath, and that one needs to put “by hand” a thermal bath connected to the dot. This was done, essentially, by putting a small imaginary part on the evolution time, which was then set to zero before taking small couplings $`\lambda ,\lambda _d`$ (that is, it was obtained a steady state where the coupling to the thermal bath is much smaller than the couplings $`\lambda ,\lambda _d`$). The most delicate question, however, concerns the fact that the steady state itself may be affected non-trivially by a thermal bath. It is probable (but this should be verified) that other, more realistic, representations of a thermal bath give the same results as rg6 in the limit of small coupling with the thermal bath. However, as the couplings $`\lambda ,\lambda _d`$ are sent to zero, due to the thermal effect of the environment, the average dot magnetization, for instance, should smoothly reach its thermal equilibrium value, and this may well be a universal cross-over behavior (since it occurs at small couplings, hence near to a second order phase transition). In order to assess this, it would be important to have a more adequate description of a thermal bath, and the theory of Caldeira and Leggett surely provides the most promising avenue. The steady-state density matrix. Our proof that the real-time perturbation theory describes a steady state was immediately adapted to the proof that the steady state can be described by a density matrix (40), almost as in an equilibrium state. As mentioned, a steady-state density matrix was also introduced in Hershfield93 for generic models under the assumption that a relaxation time was present. Our derivation is slightly different, and does not make further assumptions, since we showed that relaxation does occur in the non-equilibrium Kondo model. The main characteristics of the steady-state density matrix, as opposed to equilibrium density matrices, is that it is defined by coupling a non-local conserved charge $`Y`$ to the voltage $`V`$, instead of coupling a local conserved charge to an appropriate chemical potential. This non-locality is related to the fact that the operator $`Y`$ describes the build-up of the steady state, or, in some sense, the asymptotic state that characterizes the steady state. The properties of this asymptotic states and other consequences of this description are still to be explored. Renormalized real-time perturbation theory. We developed the two-loop renormalized real-time perturbation theory. We gave an argument for the validity of the Callan-Symanzik equation in the steady state (with the same beta functions as the equilibrium ones), and verified this to one loop for the current operator inside any correlation function. It is tempting to relate the validity of the Callan-Symanzik equation to the fact that the operator $`Y`$ appearing in the steady-state density matrix is a conserved charge. In particular, the fact that the voltage does not flow is obvious from such consideration, as it is coupled to a conserved charge. However, since $`Y`$ is non-local, it is hard to make this connection more precise. The quantum field theory gives physical quantities in the scaling limit (the universal region) $`V,T,T_KD`$, where $`T_K`$ is the Kondo temperature, which is a non-universal quantity related to the microscopic values of the couplings. Our renormalized perturbative results give the current in the region $`T_K\sqrt{V^2+T^2}D`$. This includes the part of the universal region where the system is strongly out of equilibrium. In this perturbative region, we defined a continuous family of Kondo scales $`M(\alpha )`$ depending on the ratio $`V/T=\mathrm{tan}(\alpha )`$. By comparing the scale at $`\alpha =\pi /2`$ ($`T/V0`$) with that at $`\alpha =0`$ ($`V/T0`$), we obtained a universal measure of the effect of the voltage on the Kondo cloud, as compared to the effect of the temperature. We noted that such a universal measure was correct since to two-loop order, the voltage plays the role of a good infrared cutoff for the current, so that no divergencies appear to that order as $`T/T_K0`$ (with fixed $`V/T_K`$). In connection to the latter point, it has sometimes been suggested in the literature that at $`T/T_K=0`$, the system should be in a “strong coupling regime” (see for instance ColemanHP ), and as such, the perturbation theory should not be valid and should show infrared divergencies (divergencies as $`T/T_K0`$) in higher-loop calculations. In particular, it is clear that this occurs in the calculation of any thermodynamical quantities, which can indeed be deemed “in a strong coupling regime” (at the IR fixed point) at zero temperature. In rg2 ; rg3 , for instance, one sees logarithmic divergences as $`T/T_K0`$ for any fixed $`V/T_K`$ at the one-loop order of perturbation theory for the spin susceptibility, pointing to the fact that for describing the limit $`T/T_K0`$ of that thermodynamical quantity, one needs to know about the IR fixed point<sup>5</sup><sup>5</sup>5The large-$`V/T_K`$ behavior in $`1/V`$ found at 0-loop order, then, can only be trusted in the region $`VTT_K`$. (the perturbation theory only describes the theory around its UV fixed point). Our point, though, is that this might not be so for all quantities. Our two-loop results suggest that in a sense, the current is really a dynamical quantity, ruled by the scale $`V`$. For comparison, this is much like a correlation function of two local fields is ruled by the distance between the points in equilibrium quantum field theory, even at zero temperature (the short distance behavior is described, for instance, by the UV fixed point of the theory, up to a normalization if the fields are not conserved currents, and up to the one-point functions of the operators appearing in the operator product expansion of the fields). Then, the region $`VT_K`$ should really be, for the current, a weak-coupling, UV situation, even at zero temperature; this is at least what we see at two-loop order. Notwithstanding the fact that infrared divergencies may signal that the steady state is not reached, as discussed above, possible divergencies in the perturbative expansion of the current at higher orders may correspond to simple power-like non-analyticity in the couplings at zero temperature, with smaller contributions than those of the two-loop results (as in the usual situation of correlation functions in equilibrium quantum field theory). Finally, it would be very interesting to fully verify the validity of the Callan-Symanzik equation with a magnetic field. It is easy to check for instance that the one-loop, finite-magnetic-field results of rg2 ; rg3 ; rg5 satisfy the one-loop Callan-Symanzik equations in the universal regime. Although we did not cover the case with a finite magnetic field (and see the discussion above for the subtleties involved), we expect that our general arguments for the validity of the Callan-Symanzik equations to all order still hold since in the universal regime, the magnetic field is a low-energy scale as compared to the band width. It may be useful to note that the usual perturbative renormalization was modified in rg2 ; rg3 in order to correctly incorporate the structure of logarithmic divergencies in the region $`V>D`$ (where $`D`$ is the bandwidth) of the one-loop calculation of the current and of the impurity magnetization at finite magnetic field; more precisely, energy-dependent coupling constants were introduced. We want to stress that this region may be non-universal (that is, results may depend on the precise structure of the band), as is the region $`T>D`$ or $`T_K>D`$. Then, naturally, it cannot be universally described by a finite number of coupling constants (or, more precisely, by a finite number of RG invariants). In order to recover “scaling”, one needs to use exact RG or similar methods, and the usual Callan-Symanzik equation does not hold; but this is not in disagreement with our results, which only deal with the universal regime $`V,T,T_KD`$. Acknowledgments BD is grateful to J. Cardy, J. Chalker and F. Essler, as well as all members of the QFT research group at Oxford, for sharing their insights in many occasions, to A. Lamacraft for discussions during his visit, and to N. Shah for useful comments on the manuscript. BD acknowledges support from an EPSRC post-doctoral fellowship (grant GR/S91086/01), and also acknowledges Rutgers University, where this work was initiated. NA is grateful to C. Bolech, P. Mehta, O. Parcollet, A. Rosch and A. Schiller, for numerous illuminating discussions, criticisms and suggestions as well as useful comments on the manuscript. We would like to thank S. Kehrein for various discussions during the early stage of this work. ## Appendix A Connected correlation functions Consider a theory with density matrix $`\rho `$, and denote, for any operator $`𝒪`$, $$𝒪=\frac{\mathrm{Tr}\left[\rho 𝒪\right]}{\mathrm{Tr}\left[\rho \right]}.$$ Consider the following average: $$𝒪\frac{𝒫\mathrm{exp}\left(i_{t_0}^0𝑑tH_I(t)\right)𝒪}{𝒫\mathrm{exp}\left(i_{t_0}^0𝑑tH_I(t)\right)}$$ (126) where $`H_I(t)`$ can be any operator depending on the (time) parameter $`t`$, and $`𝒪`$ is also any operator. The path-ordered exponential above is understood as an expansion in time-ordered integrals of multilinears of $`H_I(t)`$’s. The connected correlation functions, where the $`H_I(t)`$’s are “connected” to $`𝒪`$, can then naturally be defined by saying that $$𝒪=\underset{n=0}{\overset{\mathrm{}}{}}i^n_{t_0}^0𝑑t_1_{t_1}^0𝑑t_2\mathrm{}_{t_{n1}}^0𝑑t_nH_I(t_1)H_I(t_2)\mathrm{}H_I(t_n)𝒪_{\mathrm{connected}}.$$ (127) Equivalently, they can be defined recursively by $`H_I(t_1)\mathrm{}H_I(t_n)𝒪`$ $`=H_I(t_1)\mathrm{}H_I(t_n)𝒪_{\mathrm{connected}}`$ (128) $`+{\displaystyle \underset{m=1}{\overset{n}{}}}{\displaystyle \underset{\begin{array}{c}\{\alpha _1,\mathrm{},\alpha _m\}\{1,\mathrm{},n\}\\ \alpha _1<\mathrm{}<\alpha _m\\ \{\beta _1,\mathrm{},\beta _{nm}\}\{\alpha _1,\mathrm{},\alpha _m\}=\{1,\mathrm{},n\}\end{array}}{}}H_I(t_{\alpha _1})\mathrm{}H_I(t_{\alpha _m})H_I(t_{\beta _1})\mathrm{}H_I(t_{\beta _{nm}})𝒪_{\mathrm{connected}}.`$ (132) This recursive definition does not involve integrations over the time parameters $`t_i`$’s, so that it is in fact a general definition for correlation functions where a set of arbitrary operators $`H_I(t_i)`$’s parametrized by $`i=1,\mathrm{},n`$ are connected to $`𝒪`$. The main property of connected correlation functions is the following. Consider a fixed set of $`t_i`$’s for $`i=1,\mathrm{},n`$ and a fixed subset $`t_{\alpha _1},\mathrm{},t_{\alpha _m}`$ for $`m<n`$. Consider correlation functions of $`H_I(t_i)`$’s with and without insertion of $`𝒪`$. One can observe that if the operators $`H_I(t_{\alpha _i})`$’s factorize out of all such correlation functions (for instance: $`H_I(t_1)\mathrm{}H_I(t_{\alpha _1})\mathrm{}H_I(t_{\alpha _m})\mathrm{}H_I(t_n)𝒪`$ $`=H_I(t_1)\mathrm{}\widehat{H_I(t_{\alpha _1})}\mathrm{}\widehat{H_I(t_{\alpha _m})}\mathrm{}H_I(t_n)𝒪H_I(t_{\alpha _1})\mathrm{}H_I(t_{\alpha _m}))`$ then all connected correlation functions involving at least one of the $`H_I(t_{\alpha _i})`$’s are zero. This is easy to show from (A) by induction on the number of operators $`H_I(t_i)`$’s inside connected correlation functions. Connected correlation functions also occur in more general situations: $`{\displaystyle \frac{𝒫\mathrm{exp}\left(i_{t_1^i}^{t_1^f}𝑑tH_I^{(1)}(t)\right)𝒫\mathrm{exp}\left(i_{t_2^i}^{t_2^f}𝑑tH_I^{(2)}(t)\right)\mathrm{}𝒫\mathrm{exp}\left(i_{t_N^i}^{t_N^f}𝑑tH_I^{(N)}(t)\right)𝒪}{𝒫\mathrm{exp}\left(i_{t_1^i}^{t_1^f}𝑑tH_I^{(1)}(t)\right)𝒫\mathrm{exp}\left(i_{t_2^i}^{t_2^f}𝑑tH_I^{(2)}(t)\right)\mathrm{}𝒫\mathrm{exp}\left(i_{t_N^i}^{t_N^f}𝑑tH_I^{(N)}(t)\right)}}`$ $`=𝒫\mathrm{exp}\left(i{\displaystyle _{t_1^i}^{t_1^f}}𝑑tH_I^{(1)}(t)\right)𝒫\mathrm{exp}\left(i{\displaystyle _{t_2^i}^{t_2^f}}𝑑tH_I^{(2)}(t)\right)\mathrm{}𝒫\mathrm{exp}\left(i{\displaystyle _{t_N^i}^{t_N^f}}𝑑tH_I^{(N)}(t)\right)𝒪_{\mathrm{connected}}.`$ On the right-hand side, the operators $`H_I^{(1)}(t_i)`$’s, $`H_I^{(2)}(t_i)`$’s, …, $`H_I^{(N)}(t_i)`$’s are all connected to $`𝒪`$. ## Appendix B Proof of factorization Consider a product of operators of the type $$\stackrel{}{J}_1(x+x_1)\stackrel{}{S}\stackrel{}{J}_2(x+x_2)\stackrel{}{S}\mathrm{}\stackrel{}{J}_n(x+x_n)\stackrel{}{S}$$ where $`\stackrel{}{J}_{1,2,\mathrm{}}`$ can be $`\stackrel{}{J}_d,\stackrel{}{J}_x`$ or $`\stackrel{}{J}_y`$. Recall the notation (19). There, the trace is performed over the Hilbert space of the conformal field theory where the currents $`\stackrel{}{J}_{1,2,\mathrm{}}`$ act and over the two-dimensional impurity space associated to $`\stackrel{}{S}`$. Insert an operator which is composed of products of local operators acting on the CFT Hilbert space tensored with an arbitrary operator on the impurity space. We will denote it by $$\underset{a}{}𝒪^aS_a$$ where $`a=0,1,2,3`$ and $`S_0`$ is the identity $`\mathrm{𝟏}`$ on the impurity space. The operators in $`𝒪^a`$ can be at any fixed time (with respect to the theory $`H_0`$) and position, and can also be integrals of such operators over finite time intervals. We the consider the SU(2) invariant quantity: $`\stackrel{}{J}_1(x+x_1)\stackrel{}{S}\stackrel{}{J}_2(x+x_2)\stackrel{}{S}\mathrm{}\stackrel{}{J}_n(x+x_n)\stackrel{}{S}𝒪^aS_a_0`$ $`=J_1^{i_1}(x+x_1)J_2^{i_1}(x+x_2)\mathrm{}J_n^{i_n}(x+x_n)𝒪^a_0S_{i_1}S_{i_2}\mathrm{}S_{i_n}S_a_0.`$ (133) The first factor on the right-hand side factorizes as $`|x|\mathrm{}`$ because of the locality of the operators and because the correlation function is evaluated in a unitary quantum field theory: $$J_1^{i_1}(x+x_1)J_2^{i_1}(x+x_2)\mathrm{}J_n^{i_n}(x+x_n)𝒪^a_0\stackrel{|x|\mathrm{}}{}J_1^{i_1}(x_1)J_2^{i_2}(x_2)\mathrm{}J_n^{i_n}(x_n)_0𝒪^a_0$$ This field theoretic factorization induces a corresponding factorization in the spin space, as we proceed to show. The expression $`J_1^{i_1}(x_1)J_2^{i_2}(x_2)\mathrm{}J_n^{i_n}(x_n)_0`$ is a tensor in the product space of $`n`$ copies of the fundamental representation of rhe rotation group $`O(3)`$ and the only way to form a rotational invariant with it is to multiply it by a corresponding product of matrices $`S_{i_1}S_{i_2}\mathrm{}S_{i_n}`$, i.e. $`J_1^{i_1}(x_1)J_2^{i_1}(x_2)\mathrm{}J_n^{i_n}(x_n)_0S_{i_1}S_{i_2}\mathrm{}S_{i_n}\mathrm{𝟏}.`$ Hence, the product $`S_{i_1}S_{i_2}\mathrm{}S_{i_n}`$ also factorizes, and we have $$J_1^{i_1}(x_1)J_2^{i_1}(x_2)\mathrm{}J_n^{i_n}(x_n)_0𝒪^a_0S_{i_1}S_{i_2}\mathrm{}S_{i_n}_0S_a_0$$ (134) The factorization property then follows: $`\stackrel{}{J}_1(x+x_1)\stackrel{}{S}\stackrel{}{J}_2(x+x_2)\stackrel{}{S}\mathrm{}\stackrel{}{J}_n(x+x_n)\stackrel{}{S}𝒪^aS_a_0`$ $`\stackrel{|x|\mathrm{}}{}\stackrel{}{J}_1(x+x_1)\stackrel{}{S}\stackrel{}{J}_2(x+x_2)\stackrel{}{S}\mathrm{}\stackrel{}{J}_n(x+x_n)\stackrel{}{S}_0𝒪^aS_a_0.`$ (135) The operator $`H_I`$ is a linear combinations of operators of the type $`\stackrel{}{J}\stackrel{}{S}`$. The operators $`\stackrel{}{J}`$ are right-moving fields. Hence when $`H_I`$ is evolved in time with $`H_0`$ for a time $`t`$, it becomes a linear combinations of $`\stackrel{}{J}\stackrel{}{S}`$’s at position $`t`$. This shows the factorization of juxtaposed $`H_I^{(0)}(t)`$’s as $`|t|\mathrm{}`$ when they are evaluated inside a trace with operators at fixed time and position on the right. The same proof applies if such operator insertion is put on the left of juxtaposed $`H_I^{(0)}(t)`$’s. This completes the proof. ## Appendix C Callan-Symanzik equation for matrix elements In this appendix, we will justify equation (87). Consider the finite-temperature average, in the theory described by $`H|_{V=0}`$, of the regularized current operator $`𝒥_\mathrm{\Lambda }=\lambda (\stackrel{}{J}_x)_\mathrm{\Lambda }(0)\stackrel{}{S}`$ with insertions of creation and anihilation operators of the Hamiltonian: $`A_i^{}(p)`$ and $`A_i(p)`$. More precisely these are creation and annihilation operators for eigenstates of $`H|_{V=0}`$ with energy $`p`$, corresponding to the massless particles “naturally” associated to the local operator $`J_i(x)`$, where $`J_i`$ is any of the ten operators $`\stackrel{}{J}_d`$, $`\stackrel{}{J}_x`$, $`\stackrel{}{J}_y`$, $`J_z`$. Thus consider, $$\frac{\mathrm{Tr}\left(e^{\beta H|_{V=0}}A_{i_1}(p_1)\mathrm{}A_{i_m}(p_m)𝒥_\mathrm{\Lambda }A_{i_{m+1}}^{}(p_{m+1})\mathrm{}A_{i_{m+n}}^{}(p_{m+n})\right)}{\mathrm{Tr}\left(e^{\beta H|_{V=0}}\right)}.$$ (136) Since the creation and annihilation operators are eigenoperators of the Hamiltonian, the quantity (136) satisfies the Callan-Symanzik equation (83) (that is, in (83) we can replace the steady-state average of the current by this quantity). But we showed that the interacting density matrix $`e^{\beta H|_{V=0}}`$ can be obtained from the free one $`e^{\beta H_0}`$ by evolving it for an infinite time (15). In much the same way, the interacting creation and annihilation operators can be heuristically written in terms of the mode operators associated to the current algebra operators: $$A_i^{}(p)=S(0,\mathrm{})J_i^{}(p)S(\mathrm{},0),A_i(p)=S(0,\mathrm{})J_i(p)S(\mathrm{},0)$$ (where $`S(t_1,t_2)`$ is defined in (7)). Here, $`A_i`$ is any of the 10 operators $`\stackrel{}{A}_d`$, $`\stackrel{}{A}_x`$, $`\stackrel{}{A}_y`$ or $`A_z`$. Indeed, the operators $`A_i^{}(p)`$ and $`A_i(p)`$ written in this way heuristically satisfy the canonical commutation relations amongst them, and the appropriate commutation relations with Hamiltonian $`H|_{V=0}`$, written as $`S(0,\mathrm{})H_0S(\mathrm{},0)`$ (they are eigneoperators of $`H|_{V=0}`$). Hence, the quantity (136) can be written $$\frac{\mathrm{Tr}\left(e^{\beta H_0}J_{i_1}(p_1)\mathrm{}J_{i_m}(p_m)S(\mathrm{},0)𝒥_\mathrm{\Lambda }S(0,\mathrm{})J_{i_{m+1}}^{}(p_{m+1})\mathrm{}J_{i_{m+n}}^{}(p_{m+n})\right)}{\mathrm{Tr}\left(e^{\beta H_0}\right)}.$$ (137) Taking the zero-temperature limit $`\beta \mathrm{}`$, this quantity reproduces all matrix elements of the current (see eqn(86)) $`\overline{𝒥}_\mathrm{\Lambda }=S(\mathrm{},0)𝒥_\mathrm{\Lambda }S(0,\mathrm{})`$ on the Hilbert space of the free theory $`_0`$, which proves (87). In fact, to be more precise, when going more from (137) to (136), one needs to modify slightly (137). First, the mode operators $`J_i^{}(p)`$, $`J_i(p)`$ should be replaced by appropriate wave packets $`\stackrel{~}{J}_i^{}(p)`$, $`\stackrel{~}{J}_i(p)`$, obtained by integrating the associated currents $`J_i(x)`$ (times the oscillating exponential $`e^{ipx}`$) over space with a kernel vanishing (exponentially, say) as $`|x|l`$ for some wave-packet extent $`l`$. Then, the evolution operator $`S(\mathrm{},0)`$ should be replaced by $`S(t_0,0)`$ and $`S(0,\mathrm{})`$ by $`S(0,t_0^{})`$, and these two evolution operators should be put into the trace in the denominator as well, in the order in which they appear in the numerator. By using arguments involving the vanishing of connected correlation functions, it should be observed that the limits $`|t_0|l,\beta `$ and $`|t_0^{}|l,\beta `$ exist independently. Bringing both evolution operators around the density matrix $`e^{\beta H_0}`$, one should then use the steps involved in proving (15). The result, up to vanishing contributions as $`t_0,t_0^{}l,\beta `$, is (136), with operators $`A_i^{}(p)`$ and $`A_i(p)`$ replaced respectively by $$\stackrel{~}{A}_i^{}(p)=S(0,t_0^{}+i\beta )\stackrel{~}{J}_i^{}(p)S(t_0,0),\stackrel{~}{A}_i(p)=S(0,t_0^{}+i\beta )\stackrel{~}{J}_i(p)S(t_0,0).$$ These are the operators that should correspond to asymptotic states, in the limit $`|t_0|,|t_0^{}|l\beta `$ with $`t_0=t_0^{}<0`$. In this limit, one indeed recovers (137). ## Appendix D Asymptotic behavior of the function $`P(w)`$ In this appendix we evaluate the asymptotic behavior of the function $`P(w)`$ (74). The large $`w`$ behavior is easy to obtain; the first term inside the parenthesis (on the right-hand side of the second equation in (74)) disappears, and the second term $`(p1)/(e^{w(p1)}1)`$ becomes $`(p1)\mathrm{\Theta }(p<1)`$: it is non-zero only for $`p<1`$. Taking these contributions into account as well as the last term $`e^{p^2}`$, the result is: $$P(\mathrm{})=1\frac{\gamma }{2}.$$ (138) In fact, we can also obtain the next term in the large $`w`$ expansion. The next contributions can be written in the following way, by expanding the integrand: $$_0^{\mathrm{}}\frac{dp}{p}\underset{n=1}{\overset{\mathrm{}}{}}\left((p+1)e^{nw(p+1)}+\mathrm{\Theta }(p<1)(p1)e^{nw(p1)}\mathrm{\Theta }(p>1)(p1)e^{nw(1p)}\right).$$ Interchanging integration and summation (this is valid because the integration variable $`p`$ is always kept in a region where the expansion of the integrand is convergent), the integral of every term of the sum can be expressed in terms of the exponential integrals. Every term can then be expanded at large $`w`$, giving $$\underset{n=1}{\overset{\mathrm{}}{}}\left(\frac{2}{n^2w^2}\frac{12}{n^4w^4}\frac{240}{n^6w^6}\mathrm{}\right).$$ Now every term can be re-summed, and we get the large-$`w`$ asymptotic expansion (75). The small $`w`$ behavior is more subtle. The integral can be divided into two parts: $$P(w)=\left(_0^1+_1^{\mathrm{}}\right)\frac{dp}{p}\left(\frac{p+1}{e^{w(p+1)}1}\frac{p1}{e^{w(p1)}1}+e^{p^2}\right).$$ (139) The first integral can be evaluated by expanding the integrand in small $`w`$. The integrand goes as $`1+e^{p^2}+O(w)`$ at small $`w`$, so the first integral is convergent at $`w0`$, and in fact gives an expansion in Taylor series in $`w`$. In order to obtain the divergent part in $`w`$, we need only consider the second integral, and we can forget about the term $`e^{p^2}`$. Make the transformation of variable $`pp/w`$: $$_w^{\mathrm{}}\frac{dp}{p}\left(\frac{p/w+1}{e^{p+w}1}\frac{p/w1}{e^{pw}1}\right).$$ The integrand can then be expanded in $`w`$: this gives a Taylor series in $`w^2`$ starting with power 0. Each term of this Taylor series gives a convergent integral: the asymptotic behavior of each term at $`p\mathrm{}`$ is exponentially decreasing. Moreover, each term, except for the very first one, has a behavior like $`p^0`$ as $`p0`$, so that at each order in $`w^2`$, except at the zeroth order, we can evaluate the integral. We thus obtain a Taylor series in $`w^2`$. At the zeroth order, we have $$_w^{\mathrm{}}\frac{dp}{p}\left(2\frac{(p1)e^p+1}{(e^p1)^2}\right).$$ We can write it as $$_w^{\mathrm{}}\frac{dp}{p}\left(2\frac{(p1)e^p+1}{(e^p1)^2}+\mathrm{\Theta }(p<1)\right)+_w^{\mathrm{}}\frac{dp}{p}(\mathrm{\Theta }(p<1)).$$ The first integral is convergent as $`w0`$: it has again a Taylor expansion in $`w`$. The second integral is easy to evaluate, and gives $`\mathrm{ln}(w)`$. Hence, $$P(w)\mathrm{ln}(w)\text{as}w0.$$ (140) In fact, we can gather the previous missing parts to get the constant term: $$_0^1(1)𝑑p+_0^{\mathrm{}}e^{p^2}+_0^{\mathrm{}}\frac{dp}{p}\left(2\frac{(p1)e^p+1}{(e^p1)^2}+\mathrm{\Theta }(p<1)\right).$$ (141) This gives $$\frac{\sqrt{\pi }}{2}+\gamma \mathrm{ln}(2\pi )=0.374434476\mathrm{}$$ (142) so that we get (76). ## Appendix E Integrals for the two-loop calculations The integrals, as they enter in (80), are $`I_4`$ $`=`$ $`{\displaystyle \frac{3iV}{8}}{\displaystyle }dpR(p)(i\pi \delta (p)+\underset{¯}{\mathrm{P}}{\displaystyle \frac{1}{p}}){\displaystyle }dqR(q)(i\pi \delta (p+q)+\underset{¯}{\mathrm{P}}{\displaystyle \frac{1}{p+q}})\times `$ $`{\displaystyle }drR(r)(i\pi \delta (p+q+r)+\underset{¯}{\mathrm{P}}{\displaystyle \frac{1}{p+q+r}})R(pqr)\times `$ $`(2g(p+q)(f(p+q+r)+f(r))+g(q)(f(p+q+r)+f(p+r)f(q+r)+f(r))`$ $`I_5`$ $`=`$ $`{\displaystyle \frac{3iV}{8}}{\displaystyle }dpR(p)(i\pi \delta (p)+\underset{¯}{\mathrm{P}}{\displaystyle \frac{1}{p}}){\displaystyle }dqR(q)(i\pi \delta (p+q)+\underset{¯}{\mathrm{P}}{\displaystyle \frac{1}{p+q}})\times `$ $`{\displaystyle }drR(r)(i\pi \delta (p+q+r)+\underset{¯}{\mathrm{P}}{\displaystyle \frac{1}{p+q+r}})R(pqr)\times `$ $`(2g(r)(f(p+q)f(p+q+r))+g(r)(f(q)f(q+r))+g(p+r)(f(q)f(p+q+r))`$ $`{\displaystyle \frac{3iV}{4}}{\displaystyle 𝑑pR(p)^2\left(\underset{¯}{\mathrm{P}}\frac{1}{p}\right)𝑑qR(q)^2\left(i\pi \delta (p+q)+\underset{¯}{\mathrm{P}}\frac{1}{p+q}\right)\left(i\pi \delta (q)+\underset{¯}{\mathrm{P}}\frac{1}{q}\right)pf(q)}`$ $`{\displaystyle \frac{3iV}{4}}{\displaystyle 𝑑pR(p)^2\left(i\pi \delta (p)\right)𝑑qR(q)^2\left(i\pi \delta ^{}(q)+\underset{¯}{\mathrm{P}}\frac{1}{q^2}\right)pf(q)}`$ $`I_6`$ $`=`$ $`{\displaystyle \frac{3iV}{8}}{\displaystyle }dpR(p)(i\pi \delta (p)+\underset{¯}{\mathrm{P}}{\displaystyle \frac{1}{p}}){\displaystyle }dqR(q)(i\pi \delta (p+q)+\underset{¯}{\mathrm{P}}{\displaystyle \frac{1}{p+q}})\times `$ $`{\displaystyle }drR(r)(i\pi \delta (p+q+r)+\underset{¯}{\mathrm{P}}{\displaystyle \frac{1}{p+q+r}})R(pqr)\times `$ $`\left(2g(r)(f(p+q)f(p+q+r))+g(q)(f(p+q+r)+f(p+r))+4\right)`$ $`+{\displaystyle \frac{3iV}{2}}{\displaystyle 𝑑pR(p)^2\left(i\pi \delta ^{}(p)+\underset{¯}{\mathrm{P}}\frac{1}{p^2}\right)𝑑qR(q)^2\left(i\pi \delta (p+q)+\underset{¯}{\mathrm{P}}\frac{1}{p+q}\right)qg(q)}`$ $`I_7`$ $`=`$ $`{\displaystyle \frac{3iV}{8}}{\displaystyle }dpR(p)(i\pi \delta (p)+\underset{¯}{\mathrm{P}}{\displaystyle \frac{1}{p}}){\displaystyle }dqR(q)(i\pi \delta (p+q)+\underset{¯}{\mathrm{P}}{\displaystyle \frac{1}{p+q}})\times `$ (143) $`{\displaystyle }drR(r)(i\pi \delta (p+q+r)+\underset{¯}{\mathrm{P}}{\displaystyle \frac{1}{p+q+r}})R(pqr)\times `$ $`\left(2g(p+q)(f(p+q+r)+f(r))+g(p+r)(f(q)f(p+q+r))g(q+r)(f(q)+f(r))+1\right)`$ $`{\displaystyle \frac{3iV}{4}}{\displaystyle 𝑑pR(p)^2\left(\underset{¯}{\mathrm{P}}\frac{1}{p}\right)𝑑qR(q)^2\left(i\pi \delta (p+q)+\underset{¯}{\mathrm{P}}\frac{1}{p+q}\right)\left(i\pi \delta (q)+\underset{¯}{\mathrm{P}}\frac{1}{q}\right)pf(q)}`$ $`{\displaystyle \frac{3iV}{4}}{\displaystyle 𝑑pR(p)^2\left(i\pi \delta (p)\right)𝑑qR(q)^2\left(i\pi \delta ^{}(q)+\underset{¯}{\mathrm{P}}\frac{1}{q^2}\right)pf(q)}`$ where $`R(p)`$ $`=`$ $`R_\mathrm{\Lambda }(Vp)=e^{\frac{V^2p^2}{2\mathrm{\Lambda }^2}},`$ $`g(p)`$ $`=`$ $`{\displaystyle \frac{1+e^{wp}}{1e^{wp}}}.`$ (144) Recall that $`w=\beta V`$. The function $`f(p)`$ is as in (71), and is related to $`g(p)`$ by $$f(p)=(p+1)g(p+1)(p1)g(p1).$$ (145) ## Appendix F Finite contributions of the two-loop results They are given by $`[I_4+I_5+I_6]_{\mathrm{finite}}`$ $`=`$ $`6\pi V{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dq}{q}}{\displaystyle _0^q}{\displaystyle \frac{dr}{r}}\left(g_c(q+r)(f(q)+f(r))g_c(qr)(f(q)f(r))\right)`$ (146) $`{\displaystyle \frac{3\pi V}{2}}{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dq}{q}}{\displaystyle _0^q}{\displaystyle \frac{dr}{r}}\left[g_c(qr)f_c(qr)g_c(q+r)f_c(q+r)\right]`$ $`+{\displaystyle \frac{3\pi V}{2}}{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dq}{q}}{\displaystyle _0^q}{\displaystyle \frac{dr}{r}}\left[h_c(q+r){\displaystyle \frac{q}{qr}}h_c(qr)e^{r^2}\right]`$ $`+{\displaystyle \frac{3\pi V}{2}}{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dqdr}{qr}}(g_c(r)+e^{r^2})(2f(q)f(q+r)f(qr))`$ $`{\displaystyle \frac{3\pi V}{2}}{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dqdr}{qr}}(1e^{r^2})(f_c(q+r)+f_c(qr))`$ $`12\pi V{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dr}{r}}\left({\displaystyle \frac{\gamma }{2}}+\mathrm{ln}r\right)f_c(r)`$ $`{\displaystyle \frac{3\pi V}{2}}{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dr}{r}}\left({\displaystyle \frac{\gamma }{2}}+\mathrm{ln}r\right)h_c(r)`$ $`{\displaystyle \frac{3\pi V}{2}}{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dq}{q}}f_c(q)`$ $`+{\displaystyle \frac{3\pi ^3V}{2}}\left({\displaystyle \frac{3}{w}}G(w)1\right)`$ $`+3\pi V\left(u(1i)+u(1+i)u\left({\displaystyle \frac{1i}{2}}\right)u\left({\displaystyle \frac{1+i}{2}}\right)\right)`$ $`{\displaystyle \frac{\pi ^3V}{24}}`$ and $`[I_7]_{\mathrm{finite}}`$ $`=`$ $`{\displaystyle \frac{3\pi V}{2}}{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dq}{q}}{\displaystyle _0^q}{\displaystyle \frac{dr}{r}}\left[g_c(qr)(f(q)f(r))g_c(q+r)(f(q)+f(r))\right]`$ (147) $`+{\displaystyle \frac{3\pi V}{2}}{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dq}{q}}{\displaystyle _0^q}{\displaystyle \frac{dr}{r}}\left[g_c(qr)f_c(qr)g_c(q+r)f_c(q+r)\right]`$ $`{\displaystyle \frac{3\pi V}{2}}{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dq}{q}}{\displaystyle _0^q}{\displaystyle \frac{dr}{r}}\left[h_c(q+r){\displaystyle \frac{q}{qr}}h_c(qr)e^{r^2}\right]`$ $`6\pi V{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dq}{q}}{\displaystyle _0^q}{\displaystyle \frac{dr}{r}}(1e^{r^2})(1e^{q^2})f_c(q+r)`$ $`+3\pi V{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dr}{r}}\left({\displaystyle \frac{\gamma }{2}}+\mathrm{ln}r\right)f_c(r)`$ $`+{\displaystyle \frac{3\pi V}{2}}{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dr}{r}}\left({\displaystyle \frac{\gamma }{2}}+\mathrm{ln}r\right)h_c(r)`$ $`{\displaystyle \frac{3\pi V}{2}}{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dq}{q}}f_c(q)`$ $`+6\pi V\left[u(i1)+u(i1)+u\left({\displaystyle \frac{i1}{2}}\right)+u\left({\displaystyle \frac{i1}{2}}\right)\right]`$ $`+{\displaystyle \frac{3\pi V\mathrm{ln}2}{2}}+{\displaystyle \frac{5\pi ^3V}{4}}.`$ Here, we have $`g(p)`$ $`=`$ $`{\displaystyle \frac{1+e^{wp}}{1e^{wp}}}`$ $`f(p)`$ $`=`$ $`(p+1)g(p+1)(p1)g(p1)`$ $`g_c(p)`$ $`=`$ $`g(p)\mathrm{sign}(p)`$ $`f_c(p)`$ $`=`$ $`f(p)2\mathrm{s}\mathrm{i}\mathrm{g}\mathrm{n}(p)(1e^{p^2})`$ $`h_c(p)`$ $`=`$ $`f_c(p)+(1e^{p^2})g_c(p)`$ $`G(w)`$ $`=`$ $`{\displaystyle \frac{\mathrm{sinh}(w)\frac{w}{2}}{\mathrm{sinh}^2\left(\frac{w}{2}\right)}}`$ $`u(x)`$ $`=`$ $`\mathrm{dilog}(1x)`$ (148) where $`w=V/T`$.
warning/0506/nlin0506024.html
ar5iv
text
# Proof of Nishida’s conjecture on anharmonic lattices ## 1 Introduction The Fermi-Pasta-Ulam (FPU) lattice is the famous discrete model for a continuous nonlinear string, introduced by E. Fermi, J. Pasta and S. Ulam . It consists of a number of equal point masses that nonlinearly interact with their nearest neighbors. The physical variables of the FPU lattice are the positions $`q_j`$ of the particles, see Figure 1, and their conjugate momenta $`p_j`$. Fermi, Pasta and Ulam were interested in the statistical properties of the nonlinear FPU lattice. In fact, they expected that it would attain a thermal equilibrium, as was predicted by laws in statistical mechanics. This means that the initial energy of the lattice would be redistributed and, averaged over time, equipartitioned among all the Fourier modes of the lattice, see . They performed a numerical experiment to investigate how and at what time-scale this would occur. The astonishing result of their integrations was that there was no sign of energy equipartition at all, see and : energy that was initially put in one Fourier mode, was shared by only a few other modes. Moreover, within a rather short time all the energy in the system returned to the initial mode. This recurrent behaviour has been observed in experiments on the FPU lattice with quite small as well as very large numbers of particles, on short and long time-scales, and we are led to believe that at low energy the FPU lattice behaves more or less quasi-periodically. This observation was a big surprise. On the other hand, when the initial energy of the lattice is larger then a certain threshold, equipartition indeed occurs. For a theoretical understanding of the Fermi-Pasta-Ulam experiment, one has always tried to link the FPU lattice to a completely integrable system. These are dynamical systems possessing a complete set of integrals of motion and therefore they display the regular type of behavior that was observed in the FPU experiment. With this in mind, two main approaches have been proposed. Firstly, some unexpectedly regular phenomena were observed numerically by Zabuski and Kruskal in the Korteweg-de Vries (KdV) equation. It was later proved by Gardner et al. that the latter equation is integrable. In fact, Peter Lax realised that KdV is a member of a hierarchy of integrable equations that have a Lax-pair representation and therefore a complete set of integrals. See for a good overview of these results. Using the line of thought of Lax, Flashka could for instance prove the complete integrability of the so-called Toda-lattice, which is a finite-dimensional Hamiltonian system very similar to the FPU lattice. On the other hand, one may hope to construct an integrable partial differential equation describing the behavior of the FPU lattice with some accuracy. Thus one assumes the existence of a smooth interpolation function $`u:[0,1]`$, and writes $`q_j=u(j/n)`$. If we now let $`n`$ grow to infinity and choose appropriate $`n`$-dependent scalings of the Hamiltonian function of the FPU system, we can -at least formally- obtain a partial differential equation for $`u`$ and hope to find that it is an integrable one. It is important to realise though that this procedure is not so well-defined: different ways of approximating the derivatives of $`u`$ may lead to different partial differential equations for $`u`$, see also in which among others a Boussinesq equation is obtained. Moreover, it is a priori not clear if the solutions of the resulting PDE constitute good approximations for $`q`$, even on finite time intervals. Although several heuristic arguments are available that link FPU to for instance KdV, see again , I have not been able to find any proof of such a statement. The second approach differs greatly from that of the continuum approximations and is based on finite dimensional considerations -we shall pursue this approach in the rest of this paper. As is well-known , periodic and quasi-periodic motion is typical in completely integrable finite-dimensional Hamiltonian systems. Unfortunately, the finite dimensional FPU lattice does not belong to this category. One possible explanation of the recurrent behaviour of the lattice is therefore based on the famous Kolmogorov-Arnol’d-Moser (KAM) theorem , . This theorem explains that large measure Cantor sets of quasi-periodic motions can also exist in classes of nonintegrable Hamiltonian systems, namely small perturbations of certain integrable Hamiltonian systems. The only restrictive requirement is that the integrable system that we are perturbing satisfies a certain nondegeneracy condition, which requires that each quasi-periodic motion of the integrable system has a different frequency. Even though various -again heuristic- arguments advocate this approach, and I mention in particular , the big problem is that it is not at all a priori clear whether the FPU lattice can really be viewed as a perturbation of such a nondegenerate integrable Hamiltonian system. The only obvious integrable approximation to the FPU lattice is its linearisation, which is highly degenerate as its frequency map is constant. This problem was pointed out again recently in the excellent review paper by Ford and the book by Weissert , An interesting attempt to prove the applicability of the KAM theorem was made by T. Nishida , who in 1971 published a paper that considers the FPU lattice with a finite number of particles, fixed endpoints and symmetric potential energy density function (the so-called $`\beta `$-lattice). Assuming a rather strong nonresonance condition on the frequencies of this lattice, Nishida computes its so-called Birkhoff normal form and shows that this normal form constitutes a nondegenerate integrable approximation to the original lattice Hamiltonian. Thus he proves the applicability of the KAM theorem and the existence of a positive measure set of quasi-periodic motions in the nonlinear FPU lattice. But all of this is under the assumption of a nonresonance condition, which unfortunately is only satisfied in exceptional cases. The actual value of Nishida’s computation therefore remains unclear. This paper is nevertheless devoted to a full proof of what Nishida intended to show. The main result can be summarized as follows: The Fermi-Pasta-Ulam lattice with fixed endpoints and an arbitrary finite number of moving particles possesses a completely integrable finite order Birkhoff normal form, which constitutes an integrable appoximation to the original Hamiltonian function. The integrals are the linear energies of the Fourier modes. When the potential energy density function of the lattice is an even function ($`\beta `$-lattice), this integrable approximation is nondegenerate in the sense of the KAM-theorem. This proves the existence of a large-measure set of quasi-periodic motions in the low-energy domain of the $`\beta `$-lattice. The key to proving this result lies in the fact that Nishida’s nonresonance condition, which a priori seems highly necessary for computing the Birkhoff normal form, is actually obsolete. As in , , which treat the FPU lattice with periodic boundary conditions, discrete symmetries are the key to proving Nishida’s ‘conjecture’. The results of the present paper can be considered as an extension of to the lattice with fixed endpoints with a considerably simpler proof which again uses discrete symmetry together with a simple algebraic trick. I want to remark here that the results of this paper do not provide any explicit bounds on the domain of validity of the normal form approximation. In particular we have at this moment no estimates on the behaviour of this domain when $`n`$ grows to infinity. The principal interest of the result lies in the fact that, at least to my knowledge, it is the first complete proof of the very existence of quasi-periodic motion in the FPU lattice with fixed endpoints. ## 2 The lattice equations of motion As was mentioned before, the physical variables of the FPU lattice are the positions and conjugate momenta $`(q,p)=(q_1,\mathrm{},q_n;p_1,\mathrm{},p_n)`$ of the particles in the lattice, of which we assume here that there are only finitely many. These positions and momenta are elements of the $`2n`$-dimensional cotangent bundle $`T^{}^n^{2n}`$ of $`^n`$. $`T^{}^n`$ is a symplectic manifold with the canonical symplectic form $`dqdp:=_{j=1}^ndq_jdp_j`$. Given a Hamiltonian function $`H:T^{}^n`$, the Hamiltonian vector field $`X_H`$ on $`T^{}^n`$ is implicitly defined by the relation $`dqdp(X_H,)=dH`$. The integral curves of $`X_H`$ therefore are the solutions of the system of ordinary differential equations $$\dot{q}_j=\frac{H}{p_j},\dot{p}_j=\frac{H}{q_j}$$ In the case of the FPU lattice the Hamiltonian function is the sum of the kinetic energies of all the particles and the interparticle potential energies: $$H=\underset{j}{}\frac{1}{2}p_j^2+W(q_{j+1}q_j)$$ (2.1) in which $`W:`$ is traditionally a potential energy density function of the form $$W(x)=\frac{1}{2!}x^2+\frac{\alpha }{3!}x^3+\frac{\beta }{4!}x^4$$ (2.2) The parameters $`\alpha `$ and $`\beta `$ measure the nonlinearities in the forces between the particles in the lattice. The range of the summation index $`j`$ in (2.1) depends on the exact boundary conditions that we impose on the lattice. For lattices with fixed endpoints the moving particles are labeled $`j=1,\mathrm{},n`$ and the endpoint particles are kept at rest, i.e. $`q_0=q_{n+1}=0`$ for all time. When we impose periodic boundary conditions, we label the particles by elements of the cyclic group, $`j/_N`$, so that the first and the last particle are identified, that is $`q_0=q_N`$ for all time. For conveniency we have denoted the number of particles in a periodic lattice by $`N`$ (i.e. not by $`n`$), so that its phase space becomes $`T^{}^N`$. ## 3 Discrete symmetry The Hamiltonian function (2.1) of the periodic FPU lattice (i.e. summation over $`j/_N`$) has discrete symmetries of which we shall discuss some dynamical consequences. Two important symmetries of the periodic FPU lattice are the linear mappings $`R,S:T^{}^NT^{}^N`$ defined by $`R:`$ $`(q_1,q_2,\mathrm{},q_{N1},q_N;p_1,p_2,\mathrm{},p_{N1},p_N)`$ $`(q_2,q_3,\mathrm{},q_N,q_1;p_2,p_3,\mathrm{},p_N,p_1)`$ $`S:`$ $`(q_1,q_2,\mathrm{},q_{N1},q_N;p_1,p_2,\mathrm{},p_{N1},p_N)`$ $`(q_{N1},q_{N2},\mathrm{},q_1,q_N;p_{N1},p_{N2},\mathrm{},p_1,p_N)`$ It is easily checked that $`R`$ and $`S`$ are canonical transformations that leave the periodic FPU Hamiltonian (2.1) invariant, i.e. $`R^{}(dqdp)=S^{}(dqdp)=dqdp`$ and $`R^{}H(=HR)=S^{}H(=HS)=H`$. This implies that $`R^{}X_H=X_{R^{}H}=X_H`$ and $`S^{}X_H=X_{S^{}H}=X_H`$, that is $`R`$ and $`S`$ conjugate the Hamiltonian vector field $`X_H`$ to itself. This in turn implies that $`R`$ and $`S`$ commute with the time-$`t`$ flows $`e^{tX_H}`$ of $`X_H`$. Canonical diffeomorphisms with this property are called symmetries of $`H`$ and the group of symmetries of $`H`$ is denoted $`G_H`$. The subgroup $`R,S=\{\mathrm{Id},R,R^2,\mathrm{},R^{N1},S,SR,SR^2,\mathrm{},SR^{N1}\}G_H`$ is isomorphic to the $`N`$-th dihedral group, the symmetry group of the $`N`$-gon, as its elements satisfy the multiplication relations $`R^N=S^2=\mathrm{Id}`$, $`SR=R^1S`$. As $`R`$ and $`S`$ are linear mappings, the elements of $`R,S`$ actually define a representation of $`D_N`$ in $`T^{}^N`$ by symplectic mappings. For every subgroup $`GG_H`$, we define the fixed point set $$\text{Fix}G=\{(q,p)T^{}^N|P(q,p)=(q,p)PG\}$$ (3.1) Let $`(q,p)\mathrm{Fix}G`$ and $`PG`$. Then $`P(e^{tX_H}(q,p))=e^{tX_H}(P(q,p))=e^{tX_H}(q,p)`$, i.e. $`e^{tX_H}(q,p)\mathrm{Fix}G`$. Thus we see that $`\mathrm{Fix}G`$ is an invariant manifold for the flow of $`X_H`$. ###### Proposition 3.1 When $`G`$ is compact and consists of linear symplectic isomorphisms of $`T^{}^n`$, then $`\mathrm{Fix}G`$ is a symplectic manifold with the restriction to $`\mathrm{Fix}G`$ of $`dqdp`$ as symplectic form. This implies that whenever $`X_H`$ is tangent to $`\mathrm{Fix}G`$, in particular when $`H`$ is $`G`$-symmetric, $$(X_H)|_{\mathrm{Fix}G}=X_{(H|_{\mathrm{Fix}G})}$$ Proof: Clearly, $`\mathrm{Fix}G=_{PG}\mathrm{ker}(P\mathrm{Id})`$ is a subspace of $`T^{}^N`$, and hence a submanifold. It remains to be proven that for every $`(q,p)\mathrm{Fix}G`$, the restriction of $`dqdp`$ to $`T_{(q,p)}(\mathrm{Fix}G)T_{(q,p)}(T^{}^N)`$ is nondegenerate. Let us first of all identify $`T_{(q,p)}(T^{}^N)`$ by $`T^{}^N`$ and $`T_{(q,p)}(\mathrm{Fix}G)`$ by $`\mathrm{Fix}G`$ and moreover note that since $`G`$ is compact, it contains a unique left-invariant probability measure ‘$`dP`$’, called the Haar-measure of $`G`$. We can therefore define the operator $$\mathrm{av}_G:T^{}^N\mathrm{Fix}G,v_GP(v)𝑑P$$ The operator $`\mathrm{av}_G`$ is a projection of $`T^{}^N`$ onto $`\mathrm{Fix}G`$, called ‘averaging over $`G`$’. Now let $`v\mathrm{Fix}G`$ and $`wT^{}^N`$. Then one easily computes that $`(dqdp)(v,\mathrm{av}_G(w))=(dqdp)(v,{\displaystyle _G}P(w)𝑑P)={\displaystyle _G}(dqdp)(v,P(w))𝑑P`$ $`={\displaystyle _G}(dqdp)(P(v),P(w))𝑑P={\displaystyle _G}(dqdp)(v,w)𝑑P=(dqdp)(v,w)`$ where in the second equality we have used the linearity of $`dqdp`$ in its second argument, in the third equality the fact that $`v\mathrm{Fix}G`$, and in the fourth equality that every $`PG`$ is symplectic. We now observe that when $`v\mathrm{Fix}G`$ and $`(dqdp)(v,w)=0`$ for every $`w\mathrm{Fix}G`$, then $`(dqdp)(v,w)=0`$ even for every $`wT^{}^N`$. Hence $`dqdp`$, when restricted to $`\mathrm{Fix}GT_{(q,p)}(\mathrm{Fix}G)`$, is a nondegenerate anti-symmetric bilinear form. In other words, $`\mathrm{Fix}G`$ is a symplectic subspace of $`T^{}^N`$. The final statement of this proposition follows trivially from this result. $`\mathrm{}`$ Let us now look at a periodic FPU lattice with an even number $`N=2n+2`$ of particles. Then an invariant subsystem is formed by the fixed point set of the compact group $`S=\{\mathrm{Id},S\}`$: $$\text{Fix}S=\{(q,p)T^{}^N|q_j=q_{2n+2j},p_j=p_{2n+2j}j\}$$ Clearly, in $`\text{Fix}S`$, $`q_0=q_{n+1}=p_0=p_{n+1}=0`$. Thus we see that $`\mathrm{Fix}S`$ is filled with solutions $`(q_1(t),\mathrm{},q_N(t);p_1(t),\mathrm{},p_N(t))`$ for which the $`(q_1(t),\mathrm{},q_n(t);`$ $`p_1(t),`$ $`\mathrm{},p_n(t))`$ constitute the general solution curves of the FPU lattice with fixed endpoints and $`n`$ moving particles. Hence, the FPU lattice with fixed endpoints and $`n`$ particles is embedded in the periodic lattice with $`2n+2`$ particles. By Proposition 3.1, it can be described as a Hamiltonian system on $`\mathrm{Fix}S`$, which has the restriction of $`dqdp`$ as symplectic form, and is determined by the Hamiltonian function $`H|_{\mathrm{Fix}S}`$. As coordinates on $`\mathrm{Fix}S`$ one could choose $`(q_1,\mathrm{},q_n;p_1,\mathrm{},p_n)`$. ## 4 Quasi-particles Of course, the representation of $`D_N`$ on $`T^{}^N`$ is the sum of irreducible representations. It is quite natural to choose coordinates on $`T^{}^N`$ that are adapted to these irreducible representations. For the periodic lattice, we thus make the following real-valued Fourier transformation. For $`1k<\frac{N}{2}`$ define: $`Q_k=\sqrt{{\displaystyle \frac{2}{N}}}{\displaystyle \underset{j/_N}{}}\mathrm{sin}({\displaystyle \frac{2jk\pi }{N}})q_j,`$ $`P_k=\sqrt{{\displaystyle \frac{2}{N}}}{\displaystyle \underset{j/_N}{}}\mathrm{sin}({\displaystyle \frac{2jk\pi }{N}})p_j`$ $`Q_{Nk}=\sqrt{{\displaystyle \frac{2}{N}}}{\displaystyle \underset{j/_N}{}}\mathrm{cos}({\displaystyle \frac{2jk\pi }{N}})q_j,`$ $`P_{Nk}=\sqrt{{\displaystyle \frac{2}{N}}}{\displaystyle \underset{j/_N}{}}\mathrm{cos}({\displaystyle \frac{2jk\pi }{N}})p_j`$ $`Q_N={\displaystyle \frac{1}{\sqrt{N}}}{\displaystyle \underset{j/_N}{}}q_j,`$ $`P_N={\displaystyle \frac{1}{\sqrt{N}}}{\displaystyle \underset{j/_N}{}}p_j`$ and if $`N`$ is even: $$Q_{\frac{N}{2}}=\frac{1}{\sqrt{N}}\underset{j/_N}{}(1)^jq_j,P_{\frac{N}{2}}=\frac{1}{\sqrt{N}}\underset{j/_N}{}(1)^jp_j$$ The new coordinates $`(Q,P)`$ are called quasi-particles or phonons. The transformation $`(q,p)(Q,P),T^{}^NT^{}^N`$ is symplectic and one can express the Hamiltonian in terms of $`Q`$ and $`P`$. If we write for (2.1) $$H=H_2+H_3+H_4$$ where $`H_2`$ is a quadratic polynomial in $`(q,p)`$ and $`H_3`$ and $`H_4`$ cubic and quartic polynomials in $`q`$, then we find that (see , or ) $$H_2=\underset{k=1}{\overset{N}{}}\frac{1}{2}(P_k^2+\omega _k^2Q_k^2)$$ (4.1) in which for $`k=1,\mathrm{},N`$ the numbers $`\omega _k`$ are the well-known normal mode frequencies of the periodic FPU lattice: $$\omega _k:=2\mathrm{sin}(\frac{k\pi }{N})$$ This means that written down in quasi-particles, the equations of motion of the harmonic lattice $`(\alpha =\beta =0)`$ are simply the equations for $`N1`$ uncoupled harmonic oscillators and, as $`\omega _N=0`$, one free particle. In fact, the Hamiltonian system is Liouville integrable in this situation. Integrals are for instance the linear energies $$E_k:=\frac{1}{2}(P_k^2+\omega _k^2Q_k^2)$$ The FPU model is of course much more interesting when the forces between the particles are nonlinear, i.e. when $`\alpha `$ or $`\beta `$ is nonzero. The normal modes then interact in a complicated manner that is governed by the Hamiltonians $`H_r`$ $`(r=3,4)`$, which are of the form $$H_r=\underset{\theta :|\theta |=r}{}c_\theta \underset{k=1}{\overset{N1}{}}Q_k^{\theta _k}$$ (4.2) Here the $`\theta `$ are multi-indices and the $`c_\theta `$ are real coefficients. Note that for every value of $`\alpha `$ and $`\beta `$, $`H`$ is independent of $`Q_N=\frac{1}{\sqrt{N}}_jq_j`$. Hence the total momentum $`P_N=\frac{1}{\sqrt{N}}_jp_j`$ is a constant of motion and the equations for the remaining variables are completely independent of $`(Q_N,P_N)`$. It is common to set the latter coordinates equal to zero, thus remaining with a system on $`T^{}^{N1}`$ with coordinates $`(Q_1,\mathrm{},Q_{N1},P_1,\mathrm{},P_{N1})`$. As $`\omega _1,\mathrm{},\omega _{N1}>0`$, we can conclude by the Morse-Lemma or Dirichlet’s theorem , that the origin $`(Q,P)=0`$ is a dynamically stable equilibrium of this reduced system. Assume again that $`N=2n+2`$. From the definition of the quasi-particles and the definition of $`S`$, we conclude that $`S`$ acts as follows in Fourier coordinates $`S:(Q_1,\mathrm{},Q_{N1};`$ $`P_1,\mathrm{},P_{N1})`$ $`(Q_1,\mathrm{},Q_n,Q_{n+1},\mathrm{},Q_{N1};`$ $`P_1,\mathrm{},P_n,P_{n+1},\mathrm{},P_{N1})`$ So that $$\text{Fix}S=\{(Q,P)T^{}^{N1}|Q_k=P_k=0n+1kN1\}$$ which is a symplectic manifold isomorphic to $`T^{}^n`$. Using the coordinates $`(Q_1,\mathrm{},`$ $`Q_n;`$ $`P_1,\mathrm{},`$ $`P_n)`$ on $`\text{Fix}S`$, the restriction of the symplectic form $`_{j=1}^NdQ_jdP_j`$ to $`\mathrm{Fix}S`$ is simply $`_{j=1}^ndQ_jdP_j`$. By Proposition 3.1, the Hamiltonian of the fixed endpoint lattice thus is simply the restriction of the periodic FPU Hamiltonian (4.1, 4.2) to $`\text{Fix}S`$, that is $$H|_{\text{Fix}S}=\underset{k=1}{\overset{n}{}}\frac{1}{2}(P_k^2+\mathrm{\Omega }_k^2Q_k^2)+H_3(Q_1,\mathrm{},Q_n,0,\mathrm{},0)+H_4(Q_1,\mathrm{},Q_n,0,\mathrm{},0)$$ To distinguish we have used the notation $`\mathrm{\Omega }_k:=\omega _k=2\mathrm{sin}(\frac{k\pi }{2n+2})`$ $`(1kn)`$ for the linear frequencies of the fixed endpoint lattice. ## 5 The Birkhoff normal form Nishida’s idea was to study the Hamiltonian of the fixed endpoint lattice using Birkhoff normalisation, which is a way of constructing a symplectic near-identity transformation of the phase-space with the purpose of approximating the original Hamiltonian system by a simpler one. The study of this ‘Birkhoff normal form’ can lead to important conclusions about the original system. For $`r2`$, let $`_r`$ be the finite-dimensional space of homogeneous $`r`$-th degree polynomials in $`(Q,P)`$ on $`T^{}^{N1}`$ and let $`:=_{r2}_r`$. With the Poisson bracket $`\{,\}:\times `$ defined by $$\{F,G\}:=\underset{k=1}{\overset{N1}{}}\frac{F}{q_k}\frac{G}{p_k}\frac{F}{p_k}\frac{G}{q_k}$$ $``$ is a so-called graded Lie-algebra, which means that $`\{_r,_s\}_{r+s2}`$. With this definition, we have for each $`F`$, the ‘adjoint’ linear operator $$\mathrm{ad}_F:,G\{F,G\}$$ We recall the following result, a complete proof of which can be found for instance in , and . ###### Theorem 5.1 (Birkhoff normal form theorem) Let $`H=H_2+H_3+\mathrm{}`$ be a Hamiltonian on $`T^{}^{N1}`$ such that $`H_r_r`$ for each $`r`$ and $$\text{ad}_{H_2}:G\{H_2,G\},_r_r$$ is semi-simple (i.e. complex-diagonalizable) for every $`r`$. Then for every finite $`s3`$ there is an open neighbourhood $`0UT^{}^{N1}`$ and a symplectic diffeomorphism $`\mathrm{\Phi }:UT^{}^{N1}`$ with the properties that $`\mathrm{\Phi }(0)=0`$, $`D\mathrm{\Phi }(0)=\mathrm{Id}`$ and $$\mathrm{\Phi }^{}H=H_2+\overline{H}_3+\mathrm{}+\overline{H}_s+𝒪((Q,P)^{s+1})$$ where $$\mathrm{ad}_{H_2}(\overline{H}_r)=0$$ for every $`3rs`$. The transformed and truncated Hamiltonian $`\overline{H}:=H_2+\overline{H}_3+\mathrm{}+\overline{H}_s`$ is called a Birkhoff normal form of $`H`$ of order $`s`$. Idea of proof: For $`H,F`$, the curve $`t(e^{tX_F})^{}H=He^{tX_F}`$ in $``$ satisfies the linear differential equation and initial condition $$\frac{d}{dt}(e^{tX_F})^{}H=\mathrm{ad}_F((e^{tX_F})^{}H),(e^{0X_F})^{}H=H$$ This implies that $$(e^{X_F})^{}H=e^{\mathrm{ad}_F}(H)=H+\{F,H\}+\frac{1}{2}\{F,\{F,H\}\}+\mathrm{}$$ The transformation $`\mathrm{\Phi }`$ is now constructed as the composition of a sequence of time-$`1`$ flows $`e^{X_{F_r}}`$ $`(3rs)`$ of Hamiltonian vector fields $`X_{F_r}`$ with $`F_r_r`$. The idea is that we first choose $`F_3_3`$, such that $`H`$ is transformed into $$(e^{X_{F_3}})^{}H=\underset{_2}{\underset{}{H_2}}+\underset{_3}{\underset{}{H_3+\{F_3,H_2\}}}+\underset{_4_5\mathrm{}}{\underset{}{\mathrm{}}}$$ When $`\mathrm{ad}_{H_2}`$ is semi-simple, then $`_3=\mathrm{ker}\mathrm{ad}_{H_2}\mathrm{im}\mathrm{ad}_{H_2}`$ and we can decompose $`H_3=H_3^{ker}+H_3^{im}`$ for uniquely determined $`H_3^{ker}\mathrm{ker}\mathrm{ad}_{H_2}`$ and $`H_3^{im}\mathrm{im}\mathrm{ad}_{H_2}`$. If we now choose $`F_3`$ such that $`\mathrm{ad}_{H_2}(F_3)=H_3^{im}`$, which obviously is possible, then $`(e^{X_{F_3}})^{}H=H_2+\overline{H}_3+\mathrm{}`$ for $`\overline{H}_3=H_3+\{F_3,H_2\}=H_3\mathrm{ad}_{H_2}(F_3)=H_3H_3^{im}=H_3^{ker}\mathrm{ker}\mathrm{ad}_{H_2}`$, i.e. $`\mathrm{ad}_{H_2}(\overline{H}_3)=0`$. We continue by choosing $`F_4_4`$ such that $`(e^{X_{F_4}})^{}((e^{X_{F_3}})^{}H)=H_2+\overline{H}_3+\overline{H}_4+\mathrm{}`$ for which $`\mathrm{ad}_{H_2}(\overline{H}_4)=0`$, etc. After $`s2`$ steps we obtain $`\overline{H}`$ with the desired properties. $`\mathrm{}`$ The normal form $`\overline{H}`$ is usually simpler than the original $`H`$ because it Poisson commutes with the quadratic Hamiltonian $`H_2`$. This firstly means that $`H_2`$ is a constant of motion for $`\overline{H}`$ and secondly that the flow $`te^{tX_{H_2}}`$ is a continuous symmetry of $`\overline{H}`$. Also, $`H`$ and $`\overline{H}`$ are symplectically equivalent modulo a small perturbation of order $`𝒪((Q,P)^{s+1})`$. Studying $`\overline{H}`$ instead of $`H`$ thus means neglecting this perturbation term. So we make an approximation error, but this error is very small in the low energy domain, that is for small $`(Q,P)`$. With Gronwall’s lemma, precise error estimates can be made. Finally, I would like to mention the ill-known bijective correspondence between the relative equilibria of the Birkhoff normal form and the bifurcation equations for periodic solutions obtained by Lyapunov-Schmidt reduction, as is explained in . For Hamiltonian systems with symmetry, the following elegant and well-known result is often useful, see and : ###### Theorem 5.2 Let $`H=H_2+H_3+\mathrm{}`$ and $`G`$ be a group of linear symplectic symmetries of $`H`$. Then a normal form $`\overline{H}=H_2+\overline{H}_3+\mathrm{}+\overline{H}_s`$ for $`H`$ can be constructed such that also $`\overline{H}`$ is $`G`$-symmetric. This is obvious when one realizes that the ‘generating functions’ $`F_r`$ of the proof of Theorem 5.1 can be chosen $`G`$-symmetric as well. We shall also use the following result on normal forms of symmetric subsystems, which trivially follows from Proposition 3.1 and the proof of Theorem 5.2, as the transformations $`e^{X_{F_r}}`$ induced by symmetric Hamiltonian functions $`F_r`$ leave $`\text{Fix}G`$ invariant. ###### Corollary 5.3 Let $`H`$ be a Hamiltonian function with compact symmetry group $`G`$ consisting of linear symplectic mappings. Then the normal form of $`H|_{\mathrm{Fix}G}`$ is simply the restriction of the symmetric normal form $`\overline{H}`$ of $`H`$ to $`\mathrm{Fix}G`$, i.e. $$\overline{H|_{\mathrm{Fix}G}}=\overline{H}|_{\mathrm{Fix}G}$$ This corollary tells us that it is sufficient to compute the normal form of the full system to know the normal forms of its symmetric subsystems. In particular, to find the normal form of an FPU lattice with fixed endpoints, it suffices to know the normal form of the appropriate periodic lattice. Normal forms of periodic lattices have been studied elaborately in . ## 6 Nishida’s conjecture In his 1971 paper, Nishida proved the following result: ###### Theorem 6.1 (Proven by Nishida in ) Consider the FPU lattice with fixed endpoints, $`\alpha =0`$, $`\beta 0`$ and $`n`$ arbitrary. Assume moreover the fourth order nonresonance condition on the $`\mathrm{\Omega }_k=2\mathrm{sin}(\frac{k\pi }{2n+2})`$ $`(1kn)`$ requiring that $$\underset{k=1}{\overset{n}{}}(l_km_k)\mathrm{\Omega }_k0l,m\{0,1,2,\mathrm{}\}^n\text{with}\underset{k=1}{\overset{n}{}}|l_k|+|m_k|=4\text{and}\underset{k=1}{\overset{n}{}}|l_km_k|0$$ Then the quartic Birkhoff normal form $`\overline{H}=H_2+\overline{H}_4`$ of the lattice is a function of the action variables $`a_k:=E_k/\mathrm{\Omega }_k`$ $`(1kn)`$ only and is therefore integrable. Moreover it satisfies the Kolmogorov nondegeneracy condition $$det\frac{^2\overline{H}}{a_ka_k^{}}0$$ This implies that almost all low-energy solutions of the $`\beta `$-lattice with fixed endpoints are quasi-periodic and move on invariant tori. More precisely, the relative Lebesgue measure of all these tori lying inside the small ball $`\{0H\epsilon \}`$, goes to $`1`$ as $`\epsilon `$ goes to $`0`$. As we shall see later, the numbers $$\underset{k=1}{\overset{n}{}}(l_km_k)\mathrm{\Omega }_k,\mathrm{for}\underset{k=1}{\overset{n}{}}|l_k|+|m_k|=4$$ are simply the eigenvalues of $`\mathrm{ad}_{H_2}`$ on $`_4`$. Nishida’s requirement that they be nonzero except in the trivial case that $`l_k=m_k`$ for all $`k`$ thus just means that the subspace $`\mathrm{ker}\mathrm{ad}_{H_2}_4`$ in which $`\overline{H}_4`$ must lie is very low-dimensional. It must therefore be remarked here that the integrability of the normal form follows almost trivially from Nishida’s nonresonance assumption. Nishida’s article consists mainly of the explicit computation of the normal form $`\overline{H}`$ of $`H`$ under the nonresonance assumption in order to check its nondegeneracy. But unfortunately, resonances do occur, implying that Nishida’s nonresonance condition is often violated. We have for instance the relations $$\mathrm{sin}(\pi /6)+\mathrm{sin}(3\pi /14)\mathrm{sin}(\pi /14)\mathrm{sin}(5\pi /14)=0$$ $$\mathrm{sin}(\pi /6)+\mathrm{sin}(13\pi /30)\mathrm{sin}(7\pi /30)\mathrm{sin}(3\pi /10)=0$$ $$\mathrm{sin}(\pi /2)+\mathrm{sin}(\pi /10)\mathrm{sin}(\pi /6)\mathrm{sin}(3\pi /10)=0$$ which lead to a violation of Nishida’s nonresonance condition if $`n+1`$ is a multiple of $`21`$ or $`15`$. Nishida refers to an unpublished result of Izumi proving a much stronger nonresonance condition on the $`\mathrm{\Omega }_k`$ in special cases. The result states that no $``$-linear relations between the $`\mathrm{\Omega }_k`$ exist if $`n+1`$ is a prime number or a power of $`2`$. I was not able to trace back Izumi’s proof of this statement, but note that a more general result had already been obtained in 1959 by Hemmer , who actually derived an expression for the total number of independent $``$-linear relations between the $`\mathrm{\Omega }_k`$ $`(1kn)`$ in terms of Euler’s phi-function. It turns out that no $``$-linear relations exist if and only if $`n+1`$ is a prime number or a power of $`2`$. Moreover, as the above examples illustrate, resonance relations between $`4`$ eigenvalues exist for several $`n`$ and Nishida’s condition is therefore sometimes violated. In this paper we will nevertheless prove ‘Nishida’s conjecture’ that his theorem holds without having to impose any nonresonance condition. ## 7 Near-integrability Let us start with a review of some observation in for the periodic FPU lattice. First of all we note that, as the symmetry $`R`$ is symplectic, $$(R^{}\mathrm{ad}_{H_2})(G)=R^{}\{H_2,G\}=\{R^{}H_2,R^{}G\}=\{H_2,R^{}G\}=(\mathrm{ad}_{H_2}R^{})(G)$$ where we have used that $`H_2`$ is $`R`$-symmetric. From this result we read off that $`R^{}`$ and $`\mathrm{ad}_{H_2}`$ commute as linear operators $`_r_r`$. This means that they can be diagonalized simultaneously. In this is done by introducing new canonical coordinates $`(Q,P)(z,\zeta )`$ as follows. For $`1k<\frac{N}{2}`$, we define: $`z_k`$ $`={\displaystyle \frac{1}{2}}(P_{Nk}iP_k)+{\displaystyle \frac{\omega _k}{2}}(Q_k+iQ_{Nk})`$ $`z_{Nk}`$ $`={\displaystyle \frac{1}{2}}(P_{Nk}iP_k)+{\displaystyle \frac{\omega _k}{2}}(Q_k+iQ_{Nk})`$ $`\zeta _k`$ $`={\displaystyle \frac{1}{2\omega _k}}(P_kiP_{Nk}){\displaystyle \frac{1}{2}}(Q_{Nk}+iQ_k)`$ $`\zeta _{Nk}`$ $`={\displaystyle \frac{1}{2\omega _k}}(P_kiP_{Nk})+{\displaystyle \frac{1}{2}}(Q_{Nk}+iQ_k)`$ and if $`N`$ is even: $`z_{\frac{N}{2}}={\displaystyle \frac{1}{\sqrt{2}}}(Q_{\frac{N}{2}}{\displaystyle \frac{i}{2}}P_{\frac{N}{2}}),`$ $`\zeta _{\frac{N}{2}}={\displaystyle \frac{1}{\sqrt{2}}}(P_{\frac{N}{2}}2iQ_{\frac{N}{2}})`$ It is then not hard to compute that $$H_2=\underset{1k<\frac{N}{2}}{}i\omega _k(z_k\zeta _kz_{Nk}\zeta _{Nk})+i\omega _{\frac{N}{2}}z_{\frac{N}{2}}\zeta _{\frac{N}{2}}$$ which implies that if $`\mathrm{\Theta },\theta \{0,1,2,\mathrm{}\}^{N1}`$ are multi-indices, then $$\text{ad}_{H_2}:z^\mathrm{\Theta }\zeta ^\theta \nu (\mathrm{\Theta },\theta )z^\mathrm{\Theta }\zeta ^\theta $$ in which $`\nu `$ is defined as $$\nu (\mathrm{\Theta },\theta ):=\underset{1k<\frac{N}{2}}{}i\omega _k(\theta _k\theta _{Nk}\mathrm{\Theta }_k+\mathrm{\Theta }_{Nk})+i\omega _{\frac{N}{2}}(\theta _{\frac{N}{2}}\mathrm{\Theta }_{\frac{N}{2}})$$ (7.1) In other words, $`\mathrm{ad}_{H_2}`$ is diagonal with respect to the basis of $`_r`$ consisting of the monomials $`z^\mathrm{\Theta }\zeta ^\theta `$ for which $`|\mathrm{\Theta }|+|\theta |:=_{j=1}^{N1}|\mathrm{\Theta }_j|+|\theta _j|=r`$ and the corresponding eigenvalues are the $`\nu (\mathrm{\Theta },\theta )`$. In particular we observe that $`\mathrm{ad}_{H_2}`$ is semi-simple on every $`_r`$, so that Theorem 5.1 indeed applies. A $``$-linear relation in the frequencies $`\omega _k`$ is called a resonance. For this reason, the monomials $`z^\mathrm{\Theta }\zeta ^\theta `$ for which $`\nu (\mathrm{\Theta },\theta )=0`$ are called resonant monomials. They are particularly important because they are exactly the ones that are not in $`\mathrm{im}\mathrm{ad}_{H_2}`$ and thus, as is clear from the proof of Theorem 5.1, the ones that cannot be transformed away by Birkhoff normalisation. As $`\mathrm{\Omega }_k=\omega _k(1kn)`$, Nishida’s nonresonance condition is a consequence of its analogon for periodic lattices, that can be formulated as follows: ‘When $`|\mathrm{\Theta }|+|\theta |=4`$ and $`\nu (\mathrm{\Theta },\theta )=0`$ then $`\theta _{\frac{N}{2}}\mathrm{\Theta }_{\frac{N}{2}}=0`$ and $`\theta _k\theta _{Nk}\mathrm{\Theta }_k+\mathrm{\Theta }_{Nk}=0`$ for each $`1k<\frac{N}{2}`$.’ Of course, this condition is not valid either. On the other hand, one may compute, see , that the operator $`R^{}:GGR`$ acts as follows on the coordinate function $`z_k,\zeta _k`$: $`R^{}:z_ke^{\frac{2\pi ik}{N}}z_k,\zeta _ke^{\frac{2\pi ik}{N}}\zeta _k,`$ $`z_{Nk}e^{\frac{2\pi ik}{N}}z_{Nk},\zeta _{Nk}e^{\frac{2\pi ik}{N}}\zeta _{Nk},`$ $`z_{\frac{N}{2}}z_{\frac{N}{2}}\text{and}`$ $`\zeta _{\frac{N}{2}}\zeta _{\frac{N}{2}}`$ And as a result we conclude that, as promised, $`R^{}`$ acts diagonally with respect to the monomials $`z^\mathrm{\Theta }\zeta ^\theta `$ as well: $$R^{}:z^\mathrm{\Theta }\zeta ^\theta e^{\frac{2\pi i\mu (\mathrm{\Theta },\theta )}{N}}z^\mathrm{\Theta }\zeta ^\theta $$ in which $`\mu `$ is defined as: $$\mu (\mathrm{\Theta },\theta ):=\underset{1k<\frac{N}{2}}{}j(\mathrm{\Theta }_k+\mathrm{\Theta }_{Nk}\theta _k\theta _{Nk})+\frac{N}{2}(\mathrm{\Theta }_{\frac{N}{2}}\theta _{\frac{N}{2}})\text{mod}N$$ (7.2) By Theorem 5.2 we now know that the normal form of the periodic FPU Hamiltonian must be a linear combination of monomials $`z^\mathrm{\Theta }\zeta ^\theta `$ that are both resonant and symmetric, i.e. for which $`\nu (\mathrm{\Theta },\theta )=0`$ and $`\mu (\mathrm{\Theta },\theta )=0modN`$. The following theorem was proven in . The proof below is considerably simpler though. ###### Theorem 7.1 i) The set of multi-indices $`(\mathrm{\Theta },\theta )_0^{N1}`$ for which $`|\mathrm{\Theta }|+|\theta |=3,\mu (\mathrm{\Theta },\theta )=0\text{mod}N`$ and $`\nu (\mathrm{\Theta },\theta )=0`$ is empty. ii) The set of multi-indices $`(\mathrm{\Theta },\theta )_0^{N1}`$ for which $`|\mathrm{\Theta }|+|\theta |=4,\mu (\mathrm{\Theta },\theta )=0\text{mod}N`$ and $`\nu (\mathrm{\Theta },\theta )=0`$ is contained in the set defined by the relations $`\theta _k\theta _{Nk}\mathrm{\Theta }_k+\mathrm{\Theta }_{Nk}=\theta _{\frac{N}{2}}\mathrm{\Theta }_{\frac{N}{2}}=0`$. Proof: i) Suppose that $`|\mathrm{\Theta }|+|\theta |=3`$ and $`\mu (\mathrm{\Theta },\theta )=0modN`$. Then we can conclude from looking closely at (7.2) and (7.1) that there must be integers $`k,l,m0modN`$ with $`k+l+m=0modN`$ for which $`\nu (\mathrm{\Theta },\theta )=2i\mathrm{sin}(\frac{k\pi }{N})+2i\mathrm{sin}(\frac{l\pi }{N})+2i\mathrm{sin}(\frac{m\pi }{N})=2i\mathrm{sin}(\frac{k\pi }{N})+2i\mathrm{sin}(\frac{l\pi }{N})2i\mathrm{sin}(\frac{k\pi }{N}+\frac{l\pi }{N})`$. Now I learnt the following trick from Frits Beukers: write $`2i\mathrm{sin}(\frac{k\pi }{N})=x1/x`$ and $`2i\mathrm{sin}(\frac{l\pi }{N})=y1/y`$ for some $`x,y`$ on the complex unit circle. Then $`\nu (\mathrm{\Theta },\theta )=x1/x+y1/yxy+1/xy=(1x)(1y)(1xy)/xy`$. This is zero only in the trivial cases that $`x=1`$ ($`k=0modN`$), $`y=1`$ ($`l=0modN`$) or $`xy=1`$ ($`m=0modN`$). But we already knew that $`k,l,m0modN`$. The result also follows from the convexity of the sine function. ii) The proof of the second statement is similar but more remarkable, and based on the fact that $`2i\mathrm{sin}\alpha +2i\mathrm{sin}\beta +2i\mathrm{sin}\gamma 2i\mathrm{sin}(\alpha +\beta +\gamma )=x1/x+y1/y+z1/zxyz+1/xyz=(1xy)(1xz)(1yz)/xyz`$, which again is zero in trivial cases only. $`\mathrm{}`$ In spite of Theorem 7.1, resonances do exist, as was illustrated by the examples in Section 6. A full classification of third and fourth order resonance relations in the FPU eigenvalues is given in the Appendix to . Resonance relations lead to several nontrivial resonant monomials. But according to Theorem 7.1 we now know that these nontrivial resonant monomials are not $`R`$-symmetric and hence cannot occur in the normal form of the periodic FPU lattice. As a first result, we immediately see now that there are no nonzero elements of $`_3`$ that are both resonant and $`R`$-symmetric. As a result, $`\overline{H}_3=0`$ automatically. To formulate a result for $`\overline{H}_4`$, we need to define the following Hopf-variables. For $`1k<\frac{N}{2}`$, let $`a_k:={\displaystyle \frac{1}{2\omega _k}}(P_k^2+P_{Nk}^2+\omega _k^2Q_k^2+\omega _k^2Q_{Nk}^2)`$ $`,b_k:=Q_kP_{Nk}Q_{Nk}P_k`$ $`c_k:={\displaystyle \frac{1}{2\omega _k}}(P_k^2P_{Nk}^2+\omega _k^2Q_k^2\omega _k^2Q_{Nk}^2)`$ $`,d_k:={\displaystyle \frac{1}{\omega _k}}(P_kP_{Nk}+\omega _k^2Q_kQ_{Nk})`$ and if $`N`$ is even $$a_{\frac{N}{2}}:=\frac{1}{2\omega _{\frac{N}{2}}}(P_{\frac{N}{2}}^2+\omega _{\frac{N}{2}}^2Q_{\frac{N}{2}}^2)$$ Note that $`H_2`$ can be expressed as $$H_2=\underset{1k\frac{N}{2}}{}\omega _ka_k$$ We moreover observe that when $`N=2n+2`$, the identities $`a_{\frac{N}{2}}=b_k=c_k=d_k=0`$ and $`a_k=E_k/\mathrm{\Omega }_k`$ $`(1k<\frac{N}{2})`$ hold on $`\mathrm{Fix}S`$, so that our definitions agree with the definition of $`a_k`$ in Theorem 6.1. The following result was proven in for the periodic FPU lattice. The proof consists of a careful analysis of the subspace of resonant and $`R,S`$-symmetric polynomials in $`_4`$ with the help of Theorem 7.1. It is not very deep and we do not repeat it here. ###### Theorem 7.2 Let $`H=H_2+H_3+H_4`$ be the periodic FPU Hamiltonian (4.1, 4.2). Then there is a unique quartic Birkhoff normal form $`\overline{H}=H_2+\overline{H}_4`$ of $`H`$ which is $`R,S`$-symmetric. For this normal form we have $`\overline{H}_3=0`$, whereas $`\overline{H}_4`$ is a linear combination of the quartic terms $`a_ka_k^{}`$, $`b_kb_k^{}`$ $`(1k,k^{}<\frac{N}{2})`$ and if $`N`$ is even also $`a_{\frac{N}{2}}a_k`$ $`(1k\frac{N}{2})`$ and $`d_kd_{\frac{N}{2}k}c_kc_{\frac{N}{2}k}`$ $`(1k\frac{n}{4})`$. ###### Corollary 7.3 (Conjectured by Nishida in ) Independent of $`n,\alpha `$ and $`\beta `$, the quartic Birkhoff normal form $`\overline{H}=H_2+\overline{H}_4`$ of the FPU lattice with fixed endpoints (4) is integrable with integrals $`E_k`$. Proof: By Corollary 5.3, the Birkhoff normal form of (4) is the restriction of the Birkhoff normal form of the periodic lattice with $`N=2n+2`$ particles, to $`\text{Fix}S`$. But on $`\text{Fix}S`$, we have that $`b_k=c_k=d_k=0`$ and $`a_k=E_k/\mathrm{\Omega }_k`$. So according to Theorem 7.2, $`\overline{H}_4`$ is a quadratic function of the Poisson commuting $`E_k`$. So, clearly, is $`H_2`$. $`\mathrm{}`$ Note that to prove the integrability of the normal form of the fixed endpoint lattice, we had to use the symmetry of the periodic lattice in which it is embedded. It must also be remarked here that it is very exceptional for a high-dimensional resonant Hamiltonian system to have an integrable normal form. Let us dwell a little longer on the dynamics of the normal form and consider the integral map $`E:T^{}^n^n`$ that sends $`(Q,P)(E_1,\mathrm{},E_n)`$. One checks that when $`E_k>0`$ for every $`k`$, the derivatives $`DE_k(Q,P)`$ are all linearly independent. As the level sets of $`E`$ are moreover compact, the theorem of Liouville-Arnol’d ensures that for each $`e=(e_1,\mathrm{},e_n)`$ with $`e_k>0`$ for each $`k`$, the level set $`E^1(\{e\})`$ is a smooth $`n`$-dimensional torus. To compute the flow on these tori, we transform to action-angle coordinates $`(Q,P)(a,\phi )`$ as follows. Let $`\text{arg}:^2\backslash \{(0,0)\}/_{2\pi }`$ be the argument function, $`\text{arg}:(r\mathrm{cos}\mathrm{\Phi },r\mathrm{sin}\mathrm{\Phi })\mathrm{\Phi }`$ and define $$\phi _k=\text{arg}(P_k,\mathrm{\Omega }_kQ_k),a_k=E_k/\mathrm{\Omega }_k=\frac{1}{2\mathrm{\Omega }_k}(P_k^2+\mathrm{\Omega }_k^2Q_k^2),1kn$$ With the formula $`d\text{arg}(x,y)=\frac{xdyydx}{x^2+y^2}`$, one can verify that $`(\phi ,a)`$ are canonical coordinates: $`dQdP=d\phi da`$. So in these coordinates the equations of motion read $$\dot{a}_k=0,\dot{\phi }_k=\mathrm{\Omega }_k+\frac{\overline{H}_4(a)}{a_k}$$ This simply defines periodic or quasi-periodic motion. Remark: $`(\varphi ,a)`$ are sometimes called ‘symplectic polar coordinates’. ## 8 Nondegeneracy To verify that the normal form $`\overline{H}`$ is nondegenerate, we examine the frequency map $`\mathrm{\Omega }`$ which assigns to each invariant torus the frequencies of the flow on it: $$\mathrm{\Omega }:a(\mathrm{\Omega }_1+\frac{\overline{H}_4(a)}{a_1},\mathrm{},\mathrm{\Omega }_n+\frac{\overline{H}_4(a)}{a_n})$$ The nondegeneracy condition of the KAM theorem requires that $`\mathrm{\Omega }`$ be a local diffeomorphism, which is the case if and only if the constant derivative matrix $`\frac{^2\overline{H}_4}{a_ka_k^{}}`$ is invertible. To check this, we will unfortunately need to compute the Birkhoff normal form explicitly, where until now we had been able to avoid this. In the next Theorem we shall present the normal form of the FPU Hamiltonian in the case that $`H_3=0`$, i.e. $`\alpha =0`$. This lattice, that has no cubic terms, is usually referred to as the $`\beta `$-lattice. ###### Theorem 8.1 (Conjectured by Nishida in ) If $`\alpha =0`$, then a quartic Birkhoff normal form of FPU lattice with fixed endpoints is given by $`\overline{H}=H_2+\overline{H}_4`$, where $$\overline{H}_4=\frac{\beta }{2n+2}\left(\underset{1k<ln}{}\frac{\mathrm{\Omega }_k\mathrm{\Omega }_l}{4}a_ka_l+\underset{1kn}{}\frac{3\mathrm{\Omega }_k^2}{32}a_k^2\right)$$ ‘Proof’: The computation of the normal form had already been performed by Nishida who obtained exactly the above normal form, but under the assumption that resonant monomials are absent in the lattice Hamiltonian. We now know that these monomials can not occur in the Hamiltonian as they are not $`R`$-symmetric in the corresponding periodic lattice. Hence Nishida’s computation gave the correct answer. The reader can find similar computations in , and of the normal form of the $`\beta `$-lattice with periodic boundary conditions. We can therefore obtain the result also by substituting $`a_k=E_k/\mathrm{\Omega }_k`$, $`b_k=c_k=d_k=0`$ on $`\mathrm{Fix}S`$ in the normal form of the periodic lattice that was obtained for instance in Theorem 10.1 in . $`\mathrm{}`$ It is now an easy excercise to prove the invertibility of the matrix $`\frac{^2\overline{H}_4}{a_ka_k^{}}`$. Its nondegeneracy was also checked by Nishida himself by applying elementary row and column operations to compute the determinant that turns out to be nonzero. Thus we conclude: ###### Corollary 8.2 (Conjectured by Nishida in ) If $`\alpha =0`$ and $`\beta 0`$, then the integrable quartic Birkhoff normal form $`\overline{H}=H_2+\overline{H}_4`$ of the FPU lattice with fixed endpoints (4) satisfies the Kolmogorov nondegeneracy condition. Hence almost all low-energy solutions of the FPU lattice with fixed endpoints are quasi-periodic and move on invariant tori. In fact, the relative measure of all these tori lying inside the small ball $`\{0H\epsilon \}`$, goes to $`1`$ as $`\epsilon `$ goes to $`0`$. Nishida, and we, chose to compute normal form $`H_2+\overline{H}_4`$ only for the $`\beta `$-lattice. This computation is already quite long, but it becomes extremely hard when $`\alpha 0`$. It should nevertheless also be possible to write down an expression for the fixed endpoints normal form if $`\alpha 0`$. For checking Kolmogorov’s condition this will actually be necessary. We know a priori that the resulting normal form will be integrable and depends quadratically on the $`E_k`$ (or $`a_k`$). It is very likely that for a large open set of $`\alpha `$ and $`\beta `$ the nondegeneracy condition holds and the KAM theorem applies. Without computation this is clear for $`|\alpha ||\beta |`$ (and $`n`$ fixed) because then the coefficients of the normal form can differ only slightly from those in Theorem 8.1. ## 9 Acknowledgement The author would like to address many thanks to F. Verhulst, J.J. Duistermaat, G. Gallavotti and F. Beukers for many valuable remarks and comments, without which this paper would not have been written.
warning/0506/math0506015.html
ar5iv
text
# Equivariant Giambelli and determinantal restriction formulas for the Grassmannian ## 1. The set up Fix once and for all two positive integers $`d`$ and $`n`$ with $`dn`$. Let $`V`$ be an $`n`$-dimensional complex vector space, and $`𝔾_{d,n}`$ the Grassmannian of $`d`$-dimensional linear subspaces of $`V`$. The defining action of the general linear group $`GL(V)`$ on $`V`$ induces an action on $`𝔾_{d,n}`$. We are interested in the $`T`$-equivariant integral cohomology ring $``$ of $`𝔾_{d,n}`$, where $`T`$ is a fixed maximal torus of $`GL(V)`$. We refer to \[6, §2\] and the references in that paper for the details that we leave out in this section. The natural map from $`𝔾_{d,n}`$ to $`\mathrm{Spec}()`$ induces an $`𝒮`$-algebra structure on $``$, where $`𝒮:=H_T^{}(\mathrm{Spec}())`$ is the $`T`$-equivariant integral cohomology ring of $`\mathrm{Spec}()`$ (namely the ordinary integral cohomology ring of the classifying space of $`T`$). The $`𝒮`$-algebra $``$ is independent of the choice of $`T`$ because any two maximal tori in $`GL(V)`$ are conjugate. The choice of a maximal torus $`T`$ amounts to the choice of an unordered vector space basis $``$ of $`V`$: the elements of $`T`$ are precisely those invertible linear transformations for which each element of $``$ is an eigenvector. Each element $`b`$ of $``$ thus defines a character $`ϵ_b`$ of $`T`$ that sends elements of $`T`$ to their respective eigenvalues with respect to $`b`$. The collection $`\{ϵ_b|b\}`$ forms an integral basis for the group $`X(T)`$ of characters of $`T`$. The ring $`𝒮`$ is graded isomorphic to the symmetric algebra of the abelian group $`X(T)`$ with $`X(T)`$ living in degree $`2`$. We may therefore identify $`𝒮`$ with the polynomial ring $`[ϵ_b|b]`$, where the $`ϵ_b`$ are variables in degree $`2`$. Since $`𝔾_{d,n}`$ is a smooth variety on which $`T`$ acts algebraically with finitely many fixed points, it follows that $``$ is a free $`𝒮`$-module with basis the (equivariant) classes of the Schubert subvarieties. These subvarieties are defined with respect to a fixed Borel subgroup $`B`$ containing $`T`$: they are the closures of the $`B`$-orbits in $`𝔾_{d,n}`$. The formulas of this paper are independent of the choice of $`B`$ because any two such Borel subgroups are conjugate by an element in the normalizer of $`T`$ in $`GL(V)`$. The choice of a Borel subgroup $`B`$ containing $`T`$ amounts to the choice of an ordering on the elements of the basis $``$. Let $`b_1,\mathrm{},b_n`$ be the elements of $``$ thus ordered. We have $`𝒮=[ϵ_1,\mathrm{},ϵ_n]`$, where $`ϵ_j:=ϵ_{b_j}`$. There is a one-to-one correspondence between the $`B`$-orbits and the $`T`$-fixed points in $`𝔾_{d,n}`$: each $`B`$-orbit contains one and only one $`T`$-fixed point. The $`T`$-fixed points are indexed by the subsets of cardinality $`d`$ of $``$. Denote by $`I(d,n)`$ the set of subsets of cardinality $`d`$ of $`\{1,\mathrm{},n\}`$. We use $`u,v,w,\mathrm{}`$ to denote elements of $`I(d,n)`$. For $`u`$ in $`I(d,n)`$, we write $`u=(u_1,\mathrm{},u_d)`$ where $`u_1,\mathrm{},u_d`$ are the elements of $`u`$ arranged in increasing order: $`1u_1<\mathrm{}<u_dn`$. Given $`u=(u_1,\mathrm{},u_d)`$ in $`I(d,n)`$, denote by $`e^u`$ the $`T`$-fixed point of $`𝔾_{d,n}`$ that is the span of $`\{b_{u_1},\mathrm{},b_{u_d}\}`$, by $`X(u)`$ the closure of the $`B`$-orbit containing $`e^u`$, by $`[X(u)]`$ the $`T`$-equivariant class in $``$ of the Schubert subvariety $`X(u)`$, and by $`[X(u)]_{\mathrm{cl}}`$ the ordinary cohomology class of $`X(u)`$. For each $`T`$-fixed point $`e^v`$, $`v`$ in $`I(d,n)`$, we have a natural “restriction” map $`\mathrm{Res}_v::=H_T^{}(𝔾_{d,n})𝒮:=H_T^{}(e^v)`$ induced by the inclusion of $`\{e^v\}`$ in $`𝔾_{d,n}`$. The direct product of these is an injection of rings: (1) $$\mathrm{Res}_v:\underset{vI(d,n)}{}H_T^{}(e^v)$$ For $`u`$ and $`v`$ in $`I(d,n)`$, denote by $`[X(u)]|_v`$ the image in $`𝒮`$ under $`\mathrm{Res}_v`$ of the equivariant class $`[X(u)]`$. The image of $``$ under $`\mathrm{Res}_v`$ has a neat description but we will have no use for this here: a tuple $`(\alpha _v)_{vI(d,n)}`$ in $`_{I(d,n)}𝒮`$ belongs to the image of $``$ under $`\mathrm{Res}_v`$ if and only if, whenever $`w`$ and $`x`$ in $`I(d,n)`$ are so related that there exist integers $`i`$ and $`j`$, $`1i,jn`$, with $`x=(w\{j\})\{i\}`$, it holds that $`ϵ_jϵ_i`$ divides $`\alpha _x\alpha _w`$. ## 2. An equivariant Giambelli formula Given $`u=(u_1,\mathrm{},u_d)`$ in $`I(d,n)`$, set $$\lambda _1:=nd+1u_1,\mathrm{}\lambda _i:=nd+iu_i,\mathrm{}\lambda _d:=nu_d.$$ Then $`nd\lambda _1\lambda _2\mathrm{}\lambda _d0`$. If $`u`$ is such that $`\lambda _2=\mathrm{}=\lambda _d=0`$, we call the Schubert variety $`X(u)`$ and its cohomology class special; furthermore the equivariant and ordinary cohomology classes $`[X(u)]`$ and $`[X(u)]_{\mathrm{cl}}`$ are denoted instead by $`[\lambda _1]`$ and $`[\lambda _1]_{\mathrm{cl}}`$. We extend the terminology and notation to all integers by setting $`[p]:=0`$ if $`p`$ is outside the range $`0`$, $`1`$, …, $`nd`$. Observe that $`[p]`$ belongs to $`H_T^{2p}(𝔾_{d,n})`$, which explains the notation. The classical Giambelli formula gives an expression for an arbitrary Schubert class in the ordinary cohomology ring of the Grassmannian $`𝔾_{d,n}`$ as the determinant of a $`d\times d`$ matrix whose entries are special classes. For $`u=(u_1,\mathrm{},u_d)`$ in $`I(d,n)`$, we have, from \[3, Eq.(10), page146\] for example, $`[X(u)]_{\text{cl}}=`$ $$\left|\begin{array}{cccccc}[\lambda _1]_{\text{cl}}& [\lambda _1+1]_{\text{cl}}& \mathrm{}& [\lambda _1+j1]_{\text{cl}}& \mathrm{}& [\lambda _1+d1]_{\text{cl}}\\ [\lambda _21]_{\text{cl}}& [\lambda _2]_{\text{cl}}& \mathrm{}& [\lambda _2+j2]_{\text{cl}}& \mathrm{}& [\lambda _2+d2]_{\text{cl}}\\ \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}\\ [\lambda _i+1i]_{\text{cl}}& [\lambda _i+2i]_{\text{cl}}& \mathrm{}& [\lambda _i+ji]_{\text{cl}}& \mathrm{}& [\lambda _i+di]_{\text{cl}}\\ \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}\\ [\lambda _d+1d]_{\text{cl}}& [\lambda _d+2d]_{\text{cl}}& \mathrm{}& [\lambda _d+jd]_{\text{cl}}& \mathrm{}& [\lambda _d]_{\text{cl}}\end{array}\right|$$ The $`i`$th entry on the main diagonal is $`[\lambda _i]`$ and the index increases by $`1`$ per column as we move rightwards in the same row. The subscript “cl” is to remind us that the classes are in ordinary cohomology. The theorem below gives an equivariant version of the above formula. The proof of the equivariant version does not use the ordinary version of the formula. In fact, it gives another proof of the ordinary version by specialization. Let $`u=(u_1,\mathrm{},u_d)`$ in $`I(d,n)`$. For $`i`$, $`j`$ integers such that $`1i,jd`$, set (2) $$u[i,j]=\underset{k=0}{\overset{j1}{}}c(u_i,j,k)[\lambda _i+jik]$$ where $`c(u_i,j,k):=(1)^kh_k(ϵ_{u_ij+1+k},\mathrm{},ϵ_{u_i})`$—here $`h_k`$ is the “complete symmetric polynomial,” the sum of all monomials of degree $`k`$ in the elements $`ϵ_{u_ij+1+k}`$, …, $`ϵ_{u_i}`$ of $`H_T^{}(\mathrm{pt})=𝒮`$. If $`u_ij+1+k0`$, then $`c(u_i,j,k)`$ does not make sense, but this does not matter since $`\lambda _i+jiknd+1`$ and so $`[\lambda _i+jik]=0`$ by definition. ###### Theorem 1. With notation as above, given $`u=(u_1,\mathrm{},u_d)`$ in $`I(d,n)`$, the equivariant cohomology class $`[X(u)]`$ is the determinant of the $`d\times d`$ matrix whose $`(i,j)`$th entry is $`u[i,j]`$. This theorem will be deduced in §4 from the restriction formula (Theorem 2) and the injection (1). ## 3. A determinantal formula for the restriction For integers $`p`$, $`k`$, $`r`$, set $$\mu _r^k(p):=\underset{j=k,k+1,\mathrm{},r}{}ϵ(j,p)$$ where $`ϵ(j,p):=ϵ_jϵ_p`$. This is well-defined as an element of the polynomial ring $`𝒮`$ only when $`p`$, $`k`$, and $`r`$ belong to the range $`1`$, $`2`$, …, $`n`$ and $`kr`$, but it is convenient to extend the notation somewhat: if $`k=n+1`$, the product, being over an empty index set, is taken to be $`1`$. ###### Theorem 2. Given $`u=(u_1,\mathrm{},u_d)`$ and $`v=(v_1,\mathrm{},v_d)`$ belonging to $`I(d,n)`$, the restriction $`[X(u)]|_v`$ of the $`T`$-equivariant cohomology class $`[X(u)]`$ of the Schubert variety $`X(u)`$ in the Grassmannian $`𝔾_{d,n}`$ to the $`T`$-fixed point $`e^v`$ determined by $`v`$ equals (3) $$\frac{\left|\begin{array}{cccccc}\mu _n^{u_1+1}(v_1)& \mu _n^{u_1+1}(v_2)& \mathrm{}& \mu _n^{u_1+1}(v_j)& \mathrm{}& \mu _n^{u_1+1}(v_d)\\ \mu _n^{u_2+1}(v_1)& \mu _n^{u_2+1}(v_2)& \mathrm{}& \mu _n^{u_2+1}(v_j)& \mathrm{}& \mu _n^{u_2+1}(v_d)\\ \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}\\ \mu _n^{u_i+1}(v_1)& \mu _n^{u_i+1}(v_2)& \mathrm{}& \mu _n^{u_i+1}(v_j)& \mathrm{}& \mu _n^{u_i+1}(v_d)\\ \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}\\ \mu _n^{u_d+1}(v_1)& \mu _n^{u_d+1}(v_2)& \mathrm{}& \mu _n^{u_d+1}(v_j)& \mathrm{}& \mu _n^{u_d+1}(v_d)\end{array}\right|}{\left|\begin{array}{cccccc}1& 1& \mathrm{}& 1& \mathrm{}& 1\\ ϵ_{v_1}& ϵ_{v_2}& \mathrm{}& ϵ_{v_j}& \mathrm{}& ϵ_{v_d}\\ ϵ_{v_1}^2& ϵ_{v_2}^2& \mathrm{}& ϵ_{v_j}^2& \mathrm{}& ϵ_{v_d}^2\\ \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}\\ ϵ_{v_1}^i& ϵ_{v_2}^i& \mathrm{}& ϵ_{v_j}^i& \mathrm{}& ϵ_{v_d}^i\\ \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}\\ ϵ_{v_1}^{d1}& ϵ_{v_2}^{d1}& \mathrm{}& ϵ_{v_j}^{d1}& \mathrm{}& ϵ_{v_d}^{d1}\end{array}\right|}$$ The denominator in the above expression for $`[X(u)]|_v`$ is the Vandermonde determinant which equals $$ϵ(v_2,v_1)\left(ϵ(v_3,v_1)ϵ(v_3,v_2)\right)\mathrm{}\left(ϵ(v_d,v_1)\mathrm{}ϵ(v_d,v_{d1})\right)$$ The proof of this theorem occupies sections 5 and 6. ## 4. Proof of the equivariant Giambelli In this section, Theorem 1 is deduced from Theorem 2. Because of the injection (1), it is enough to show that, for an arbitrary $`v=(v_1,\mathrm{},v_d)`$ in $`I(d,n)`$, the restriction $`[X(u)]|_v`$ is the determinant of the $`d\times d`$ matrix whose $`(i,j)\mathrm{th}`$ entry is $`u[i,j]|_v`$. We first obtain a determinantal formula for $`u[i,j]|_v`$: (4) $$u[i,j]|_v=det(N)/𝔙(v),$$ where $$𝔙(v):=(ϵ_{v_2}ϵ_{v_1})(ϵ_{v_3}ϵ_{v_1})(ϵ_{v_3}ϵ_{v_2})\mathrm{}(ϵ_{v_d}ϵ_{v_1})\mathrm{}(ϵ_{v_d}ϵ_{v_{d1}})$$ ($`𝔙`$ stands for Vandermonde) and $`N`$ denotes the following matrix (see §2 for definition of $`\mu _r^k(p)`$) $$\left(\begin{array}{ccccc}\mu _n^{u_i+1}(v_1)(ϵ_{v_1})^{j1}& \mathrm{}& \mu _n^{u_i+1}(v_s)(ϵ_{v_s})^{j1}& \mathrm{}& \mu _n^{u_i+1}(v_d)(ϵ_{v_d})^{j1}\\ \mu _n^{nd+3}(v_1)& \mathrm{}& \mu _n^{nd+3}(v_s)& \mathrm{}& \mu _n^{nd+3}(v_d)\\ \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}\\ \mu _n^{nd+r+1}(v_1)& \mathrm{}& \mu _n^{nd+r+1}(v_s)& \mathrm{}& \mu _n^{nd+r+1}(v_d)\\ \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}\\ \mu _n^n(v_1)& \mathrm{}& \mu _n^n(v_s)& \mathrm{}& \mu _n^n(v_d)\\ 1& \mathrm{}& 1& \mathrm{}& 1\end{array}\right)$$ To prove (4), we substitute for the restrictions of the special classes on the right side of (2) the determinantal expressions given by Theorem 2: $$\begin{array}{ccc}\hfill u[i,j]|_v& =& \underset{k=0}{\overset{j1}{}}c(u_i,j,k)[\lambda _i+jik]|_v\hfill \\ & =& \underset{k=0}{\overset{j1}{}}c(u_i,j,k)\frac{det(𝔑𝔲𝔪(\lambda _i+jik))}{𝔙(v)}\hfill \end{array}$$ where we have written $`𝔑𝔲𝔪(\lambda _i+jik)`$ for the $`d\times d`$ matrix whose determinant is the numerator of the expression for $`[\lambda _i+jik]|_v`$ given by Theorem 2. Rows $`2`$ through $`d`$ of $`𝔑𝔲𝔪(\lambda _i+jik)`$ do not change as $`k`$ varies: they are the same as the corresponding ones of the matrix $`N`$ in (4). And the first row of $`𝔑𝔲𝔪(\lambda _i+jik)`$ is $$(\mu _n^{u_ij+k+2}(v_1),\mathrm{},\mu _n^{u_ij+k+2}(v_d)).$$ So (4) follows once we prove (5) $`{\displaystyle \underset{k=0}{\overset{j1}{}}}c(u_i,j,k)\mu _n^{u_ij+k+2}(v_s)`$ $`=`$ $`(ϵ_{v_s})^{j1}\mu _n^{u_i+1}(v_s).`$ The above identity is the special case $`m=j1`$ of the following more general identity: for $`0mj1`$, (8) $`\begin{array}{c}{\displaystyle \underset{k=0}{\overset{m}{}}}(1)^kh_k(ϵ_{u_ij+1+k},\mathrm{},ϵ_{u_i})\mu _n^{u_ij+k+2}(v_s)=\hfill \\ (1)^mh_m(ϵ_{v_s},ϵ_{u_ij+2+m},\mathrm{},ϵ_{u_i})\mu _n^{u_ij+m+2}(v_s)\hfill \end{array}`$ The proof of (8) is by induction on $`m`$. First note that it holds for $`m=0`$. Now for the induction step: assuming that it is true for $`m`$, we show it holds for $`m+1`$. Taking $`(1)^{m+1}\mu _n^{u_ij+m+3}(v_s)`$ common out of the two terms in the following $$\begin{array}{c}(1)^mh_m(ϵ_{v_s},ϵ_{u_ij+2+m},\mathrm{},ϵ_{u_i})\mu _n^{u_ij+m+2}(v_s)+\hfill \\ (1)^{m+1}h_{m+1}(ϵ_{u_ij+2+m},\mathrm{},ϵ_{u_i})\mu _n^{u_ij+m+3}(v_s)\hfill \end{array}$$ we need only show that $$\begin{array}{c}h_m(ϵ_{v_s},ϵ_{u_ij+2+m},\mathrm{},ϵ_{u_i})(ϵ_{u_ij+m+2}ϵ_{v_s})\hfill \\ +h_{m+1}(ϵ_{u_ij+2+m},\mathrm{},ϵ_{u_i})\hfill \\ =h_{m+1}(ϵ_{v_s},ϵ_{u_ij+3+m},\mathrm{},ϵ_{u_i})\hfill \end{array}$$ but this is just the sum of the following two elementary equalities: $$\begin{array}{ccc}\hfill h_{m+1}(ϵ_{v_s},ϵ_{u_ij+3+m},\mathrm{},ϵ_{u_i})& =& h_{m+1}(ϵ_{v_s},ϵ_{u_ij+2+m},\mathrm{},ϵ_{u_i})\hfill \\ & & ϵ_{u_ij+2+m}h_m(ϵ_{v_s},ϵ_{u_ij+2+m},\mathrm{},ϵ_{u_i})\hfill \\ \hfill h_{m+1}(ϵ_{v_s},ϵ_{u_ij+2+m},\mathrm{},ϵ_{u_i})& =& h_{m+1}(ϵ_{u_ij+2+m},\mathrm{},ϵ_{u_i})\hfill \\ & & +ϵ_{v_s}h_m(ϵ_{v_s},ϵ_{u_ij+2+m},\mathrm{},ϵ_{u_i}).\hfill \end{array}$$ Now that (4) is proved, we proceed with the proof of the theorem. If we delete the $`1`$st row and $`s`$th column of $`N`$, the determinant of the resulting sub-matrix is $$𝔙(v_1,\mathrm{},\widehat{v}_s,\mathrm{},v_d):=\frac{𝔙(v)}{ϵ(v_s,v_1)\mathrm{}ϵ(v_s,v_{s1})ϵ(v_{s+1},v_s)\mathrm{}ϵ(v_d,v_s)}.$$ This follows since the determinant has degree $`0+1+\mathrm{}+d2`$ in the epsilons and is divisible by $`ϵ(v_j,v_i)`$ for $`1i,jn`$, $`i,js`$. For the determinant of $`N`$, expanding by the first row, we thus obtain $$det(N)=\underset{s=1}{\overset{d}{}}\mu _n^{u_i+1}(v_s)ϵ_{v_s}^{j1}𝔙(v_1,\mathrm{},\widehat{v}_s,\mathrm{},v_d)$$ Observe that the right side is the product of the row matrix $$(\mu _n^{u_i+1}(v_1),\mathrm{},\mu _n^{u_i+1}(v_d))$$ with the column matrix whose transpose is $$(ϵ_{v_1}^{j1}𝔙(\widehat{v}_1,v_2,\mathrm{},v_d),\mathrm{},ϵ_{v_s}^{j1}𝔙(v_1,\mathrm{},v_{d1},\widehat{v}_d))$$ This means the following for the matrix—let us call it $`M`$—whose $`(i,j)`$th entry is $`u[i,j]|_v`$: $`𝔙(v)M`$ equals $$\left(\begin{array}{ccc}\mu _n^{u_1+1}(v_1)& \mathrm{}& \mu _n^{u_1+1}(v_d)\\ \mathrm{}& \mathrm{}& \mathrm{}\\ \mu _n^{u_d+1}(v_1)& \mathrm{}& \mu _n^{u_d+1}(v_d)\end{array}\right)\left(\begin{array}{ccc}ϵ_{v_1}^0𝔙(\widehat{v}_1,\mathrm{},v_d)& \mathrm{}& ϵ_{v_1}^{d1}𝔙(\widehat{v}_1,\mathrm{},v_d)\\ \mathrm{}& \mathrm{}& \mathrm{}\\ ϵ_{v_s}^0𝔙(v_1,\mathrm{},\widehat{v}_d)& \mathrm{}& ϵ_{v_s}^{d1}𝔙(v_1,\mathrm{},\widehat{v}_d)\end{array}\right)$$ Since $`_{s=1}^d𝔙(v_1,\mathrm{},\widehat{v}_s,\mathrm{},v_d)=𝔙(v)^{d2}`$, the determinant of the matrix on the right above is $`𝔙(v)^{d1}`$. The matrix on the left—let us call it $`P`$—is the numerator in the formula for $`[X(u)]|_v`$ of Theorem 2. Taking determinants, we get $$𝔙(v)^ddet(M)=det(P)𝔙(v)^{d1}$$ and so by Theorem 2 $`det(M)=det(P)/𝔙(v)=[X(u)]|_v.`$ $`\mathrm{}`$ ## 5. Proof of the restriction formula In this section, Theorem 2 is proved. Theorem 3, which is stated and proved in §6, allows us to reduce the proof to combinatorics. More precisely, Theorem 3 tells us that if we have a Gröbner degeneration of an open piece of the Schubert variety $`X(u)`$ around the $`T`$-fixed point $`e^v`$, then we can compute the desired restriction $`[X(u)]|_v`$. Such a Gröbner degeneration is described in —indeed it was the goal of that paper to describe such a degeneration. We now recall this description. We identify $`𝔾_{d,n}`$ as the orbit space for the action on $`n\times d`$ matrices of rank $`d`$ by the group of invertible $`d\times d`$ matrices by multiplication on the right. The subset consisting of those matrices in which the submatrix determined by the rows $`v_1,\mathrm{},v_d`$ is the identity matrix gives us an affine $`T`$-stable patch of $`𝔾_{d,n}`$ around the point $`e^v`$. This patch is an affine space which we denote $`𝔸^v`$. The coordinate function $`X(r,j)`$ on $`𝔸^v`$ determined by the entry of the matrix in position $`(r,j)`$, $`rv`$, is an eigenvector for $`T`$ with character $`(ϵ_rϵ_{v_j})`$. Thus a natural way to index these coordinates on $`𝔸^v`$ is by the pairs $`(r,c)`$, $`1r,cn`$, such that $`cv`$ and $`r\{1,\mathrm{},n\}v`$—instead of $`X(r,j)`$ we write $`X(r,v_j)`$. Denote by $`^v`$ the set of all such pairs. The intersection $`Y(u)`$ of $`X(u)`$ with the affine patch $`𝔸^v`$ of $`𝔾_{d,n}`$ around $`e^v`$ is of course a closed subvariety in $`𝔸^v`$. As proved in \[7, §5\], there exist term orders on the monomials in the variables $`X(r,c)`$ with respect to which the initial ideal of the ideal of functions vanishing on $`Y(u)`$ is the face ideal of a certain simplicial complex $`𝒞_u^v`$ with vertex set $`^v`$. We want to describe the maximal faces of this complex and thereby the complex itself. But before we do that, a digression is necessary. In order that specializations to degenerate situations work smoothly, the correct definition of simplicial complex needs to be adopted. We do not insist, unlike in \[12, Chapter II\] and like in \[11, Definition 1.4\], on the axiom that singleton subsets are faces. More precisely, here are our definitions: A simplicial complex is a pair $`(V,F)`$ of a set $`V`$ and a set $`F`$ of subsets of $`V`$; the elements of $`V`$ are called vertices and those of $`F`$ faces; the following axioms hold: (1) the empty subset of $`V`$ is a face, and (2) a subset of a face is a face. Because of (2) we may replace (1) by (1’): $`F`$ is non-empty. Given a simplicial complex $`(V,F)`$, its (Stanley-Reisner) face ring $`R`$ is defined as follows: consider the polynomial ring, over some implicit base, in a set of variables indexed by $`V`$—we abuse notation and let $`V`$ itself denote the set of variables; the linear span of monomials whose support is not contained in any face forms an ideal—let us call it the face ideal (or should it be the non-face ideal?); the quotient of the polynomial ring by the face ideal is $`R`$. It is readily seen that the face ideal is the intersection, over all maximal faces, of the ideal generated by the variables in the complement of that face. The digression being over, we now start on the description of the simplicial complex $`𝒞_u^v`$. Denote by $`𝔑^v`$ the subset of $`^v`$ consisting of those pairs $`(r,c)`$ for which $`r>c`$. The element $`u`$ of $`I(d,n)`$ determines as follows a subset $`𝔖_u^v`$ of $`𝔑^v`$ with the following property: writing $`𝔖_u^v=\{(r_1,c_1),\mathrm{},(r_k,c_k)\}`$, we have $`u=\left(v\{c_1,\mathrm{},c_k\}\right)\{r_1,\mathrm{},r_k\}`$. To define $`𝔖_u^v`$, proceed by induction on $`d`$. Let $`i`$, $`1id`$, be the largest such that $`v_iu_1`$. Set $`v^{}=v\{v_i\}`$ and $`u^{}=u\{u_1\}`$. Then $`v^{}u^{}`$ and $`𝔖_u^{}^v^{}`$ is defined by induction. Set $$𝔖_u^v=\{\begin{array}{cc}𝔖_u^{}^v^{}\{(u_1,v_i)\}\hfill & \text{if }u_1v_i\hfill \\ 𝔖_u^{}^v^{}\hfill & \text{if }u_1=v_i\hfill \end{array}$$ We draw—see Example 1 and Figure 1 below—a grid with the elements of $`𝔑^v`$ being the lattice points—in the notation $`(r,c)`$, the $`r`$ is suggestive of row index and $`c`$ of column index. The solid dots in the figure denote the points of $`𝔖_u^v`$. From each solid dot $`\beta `$ we draw a vertical line and a horizontal line. Let $`\beta (\text{start})`$ and $`\beta (\text{finish})`$ denote respectively the points where the vertical and the horizontal lines meet the boundary. In Figure 1 for example $`\beta (\text{start})=(14,11)`$ and $`\beta (\text{finish})=(16,13)`$ for $`\beta =(16,11)`$. A lattice path between a pair of such points $`\beta (\text{start})`$ and $`\beta (\text{finish})`$ is a sequence $`\alpha _1,\mathrm{},\alpha _q`$ of elements of $`𝔑^v`$ with $`\alpha _1=\beta (\text{start})`$, $`\alpha _q=\beta (\text{finish})`$, and for $`j`$, $`1jq1`$, writing $`\alpha _j=(r,c)`$, $`\alpha _{j+1}`$ is either $`(r^{},c)`$ or $`(r,c^{})`$, where $`r^{}`$ (respectively $`c^{}`$) is the smallest integer not in $`v`$ (respectively in $`v`$) and greater than $`r`$ (respectively $`c`$). If $`\beta (\text{start})=(r,c)`$ and $`\beta (\text{finish})=(R,C)`$, then $`q=(Rr)+(Cc)+1`$. Let us write $`𝔖_u^v=\{\beta _1,\mathrm{},\beta _p\}`$. Consider the set of all $`p`$-tuples of paths $`\mathrm{\Lambda }=(\mathrm{\Lambda }_1,\mathrm{},\mathrm{\Lambda }_p)`$, where $`\mathrm{\Lambda }_j`$ is a lattice path between $`\beta _j(\text{start})`$ and $`\beta _j(\text{finish})`$, and no two $`\mathrm{\Lambda }_j`$ intersect. A particular such $`p`$-tuple is shown in Figure 1. Such $`p`$-tuples form an indexing set for the maximal faces of the simplicial complex $`𝒞_u^v`$: to the $`p`$-tuple $`\mathrm{\Lambda }=(\mathrm{\Lambda }_1,\mathrm{},\mathrm{\Lambda }_p)`$ corresponds the maximal face $`\mathrm{\Lambda }_1\mathrm{}\mathrm{\Lambda }_p\left(^v𝔑^v\right)`$: in the degenerate case when $`𝔖_u^v`$ is empty (which happens only if $`u=v`$), there is a unique maximal face, namely $`^v𝔑^v`$. ###### Example 1. Let $`d=14`$, $`n=27`$, $`v`$ $`=`$ $`(1,2,3,4,5,10,11,12,13,18,19,20,21,22),\text{ and}`$ $`u`$ $`=`$ $`(1,4,5,9,12,13,16,17,19,22,24,25,26,27),\text{ so that}`$ $`𝔖_u^v`$ $`=`$ $`\{(9,3),(16,11),(17,10),(24,21),(25,20),(26,18),(27,2)\}.`$ Figure 1 shows a particular tuple of non-intersecting lattice paths. $`\mathrm{}`$ Consider now the subvariety of $`𝔸^v`$ defined by the face ideal of the complex $`𝒞_u^v`$. It is a union of coordinate planes. There is one plane for each maximal face and it is defined by the vanishing of the coordinates corresponding to the vertices in the complement of that face. For a maximal face $`f`$ corresponding to $`(\mathrm{\Lambda }_1,\mathrm{},\mathrm{\Lambda }_p)`$, denote by $`m_f`$ the product, over all $`(r,c)`$ in $`𝔑^v\left(\mathrm{\Lambda }_1\mathrm{}\mathrm{\Lambda }_p\right)`$, of $`ϵ(r,c):=ϵ_rϵ_c`$. It follows from Theorem 3 that the restriction $`[X_u]|_v`$ is the sum $`m_f`$ as $`f`$ varies over all maximal faces. The last assertion holds also in the degenerate case $`u=v`$: then $`𝔖_u^v`$ is empty; $`𝒞_u^v`$ has only one maximal face, namely $`^v𝔑^v`$; and $`m_f`$ is the product over $`(r,c)`$ in $`𝔑^v`$ of $`(ϵ_rϵ_c)`$. In particular, * If $`u=v=(1,2,\mathrm{},d)`$, then $`𝒞_u^v`$ has only the empty face, and $`m_f`$ is the product over $`(r,c)`$ in $`𝔑^v=^v`$ of $`(ϵ_rϵ_c)`$. * If $`u=v=(nd+1,\mathrm{},n)`$, then the unique maximal face of $`𝒞_u^v`$ is $`f=^v`$ and $`m_f`$, being the product over an empty index set, equals $`1`$. ###### Example 2. This example is simple enough so we can easily draw all possible tuples of non-intersecting lattice paths. Let $`d=6`$, $`n=13`$, $$\begin{array}{c}v=(1,2,3,8,9,10),\text{ and }u=(4,6,7,10,11,13).\\ \text{Then }𝔖_u^v=\{(4,3),(6,2),(7,1),(11,9),(13,8)\}.\end{array}$$ Figure 2 shows all the $`5`$-tuples of non-intersecting lattice paths—there are $`9`$ of them. Writing $`ϵ(r,c)`$ for $`ϵ_rϵ_c`$, $$\begin{array}{ccc}\hfill [X_u]|_v& =& ϵ(11,1)ϵ(11,2)ϵ(11,3)ϵ(12,1)ϵ(12,2)ϵ(12,3)ϵ(13,1)ϵ(13,2)ϵ(13,3)\hfill \\ & & [ϵ(12,9)ϵ(12,10)+ϵ(13,8)ϵ(12,10)+ϵ(13,8)ϵ(13,9)]\hfill \\ & & \left[ϵ(5,3)+ϵ(6,2)+ϵ(7,1)\right].\mathrm{}\hfill \end{array}$$ Thus the proof of the restriction formula is reduced to the combinatorial problem of establishing (9) $$m_f=E(u,v)$$ where $`E(u,v)`$ stands for the expression (3). Whether this problem admits of an elegant solution by means of a Lindstrom-Gessel-Viennot type argument we do not know. What follows is an elementary argument based on induction. Proceed by induction on $`d`$. The case $`d=1`$ being easily verified, let $`d2`$. The strategy of the proof is this. In the first part, we work with $`m_f`$ and express it in terms of “smaller” $`m_f`$—those attached to simplicial complexes $`𝒞_u^{}^v^{}`$ for elements $`u^{}v^{}`$ in $`I(d1,n)`$. By the induction hypothesis, equality (9) applies to these smaller $`m_f`$, so that we get an expression for $`m_f`$ in terms of $`E(u^{},v^{})`$—the precise expression is in (12) below. In the second part, we will algebraically manipulate the expression (3) for $`E(u,v)`$ to express it in terms of $`E(u^{},v^{})`$. The resulting expression will turn out to be the same as that for $`m_f`$ obtained in the first part. This will finish the proof. So first consider $`m_f`$. Let $`r`$ be the integer, $`1rd`$, such that $`u_{r1}<v_du_r`$. Write as before $`𝔖_u^v=\{\beta _1,\mathrm{},\beta _p\}`$. It is easy to see that $$\begin{array}{c}\beta _p(\text{finish})=(u_d,v_d),\beta _{p1}(\text{finish})=(u_{d1},v_d),\mathrm{},\\ \beta _{pd+r+1}(\text{finish})=(u_{r+1},v_d);\\ \text{furthermore, }\beta _{pd+r}(\text{finish})=(u_r,v_d)\text{ unless }u_r=v_d\text{.}\end{array}$$ Figure 3 depicts the situation. Partition the set $`S`$ of the $`p`$-tuples of paths $`(\mathrm{\Lambda }_1,\mathrm{},\mathrm{\Lambda }_p)`$ (those indexing the maximal faces of $`𝒞_u^v`$) into subsets $`S_\underset{¯}{\mathrm{}}`$ indexed by sequences $`\underset{¯}{\mathrm{}}=(\mathrm{}_{r+1},\mathrm{},\mathrm{}_d)`$ of integers such that $`u_r<\mathrm{}_{r+1}u_{r+1}`$, $`u_{r+1}<\mathrm{}_{r+2}u_{r+2}`$, …, $`u_{d1}<\mathrm{}_du_d`$: the subset $`S_\underset{¯}{\mathrm{}}`$ consists of all those in which the segment joining $`(\mathrm{}_j,v_{d1})`$ and $`(\mathrm{}_j,v_d)`$ is part of the path $`\mathrm{\Lambda }_{pd+j}`$ for every $`j`$, $`r+1jd`$. That the $`S_\underset{¯}{\mathrm{}}`$ form a partition of $`S`$ is readily seen. Letting $`S_\underset{¯}{\mathrm{}}`$ also denote the corresponding partition of the maximal faces of $`𝒞_u^v`$, we obtain (10) $$m_f=\underset{\underset{¯}{\mathrm{}}}{}\underset{fS_\underset{¯}{\mathrm{}}}{}m_f$$ For the moment, let us fix a certain $`\underset{¯}{\mathrm{}}`$. Set $`u^{}:=(u_1,\mathrm{},u_{r1},l_{r+1},\mathrm{},l_d)`$ and $`v^{}:=(v_1,\mathrm{},v_{d1})`$ (although $`u^{}`$ depends on $`\underset{¯}{\mathrm{}}`$, we still write only $`u^{}`$ and not $`u^{}(\underset{¯}{\mathrm{}})`$). Then $`u^{}v^{}`$. We want to use the induction hypothesis to express $`_{fS_\underset{¯}{\mathrm{}}}m_f`$ in terms of $`E(u^{},v^{})`$. Towards this, we make two observations. First, the factor $$\mu _n^{u_d+1}(v_d)\underset{ri<d}{}\mu _{\mathrm{}_{i+1}1}^{u_i+1}(v_d)$$ is common to all the terms in $`_{fS_\underset{¯}{\mathrm{}}}m_f`$. Second, the integer $`v_d`$ does not occur as a row or column index if we restrict attention to the first $`d1`$ columns of Figure 3 (which tells us that $`E(u^{},v^{})`$ needs to be adjusted to take care of this). Consider $`E(u^{},v^{})`$—this is the expression (3) with $`u`$ and $`v`$ replaced respectively by $`u^{}`$ and $`v^{}`$. In the matrix whose determinant is the numerator of $`E(u^{},v^{})`$, the entry in position $`(i,j)`$ where $`ir1`$ has $`ϵ(v_d,v_j)`$ occurring as a factor—this factor does not occur if $`ir`$. Denote by $`E(u^{},v^{};v_d)`$ the modified expression where the factors $`ϵ(v_d,v_j)`$ are taken out—more precisely, $`E(u^{},v^{};v_d):=`$ (11) $$\frac{\left|\begin{array}{ccccc}{\displaystyle \frac{\mu _n^{u_1+1}(v_1)}{ϵ(v_d,v_1)}}& \mathrm{}& {\displaystyle \frac{\mu _n^{u_1+1}(v_j)}{ϵ(v_d,v_j)}}& \mathrm{}& {\displaystyle \frac{\mu _n^{u_1+1}(v_{d1})}{ϵ(v_d,v_{d1})}}\\ \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}\\ {\displaystyle \frac{\mu _n^{u_{r1}+1}(v_1)}{ϵ(v_d,v_1)}}& \mathrm{}& {\displaystyle \frac{\mu _n^{u_{r1}+1}(v_j)}{ϵ(v_d,v_j)}}& \mathrm{}& {\displaystyle \frac{\mu _n^{u_{r1}+1}(v_{d1})}{ϵ(v_d,v_{d1})}}\\ \mu _n^{l_{r+1}+1}(v_1)& \mathrm{}& \mu _n^{l_{r+1}+1}(v_j)& \mathrm{}& \mu _n^{l_{r+1}+1}(v_{d1})\\ \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}\\ \mu _n^{l_d+1}(v_1)& \mathrm{}& \mu _n^{l_d+1}(v_j)& \mathrm{}& \mu _n^{l_d+1}(v_{d1})\end{array}\right|}{ϵ(v_2,v_1)\left(ϵ(v_3,v_1)ϵ(v_3,v_2)\right)\mathrm{}\left(ϵ(v_{d1},v_1)\mathrm{}ϵ(v_{d1},v_{d2})\right)}$$ Looking at Figure 3 and using the induction hypothesis, we get (12) $$\underset{fS_\underset{¯}{\mathrm{}}}{}m_f=E(u^{},v^{};v_d)\mu _n^{u_d+1}(v_d)\underset{ri<d}{}\mu _{\mathrm{}_{i+1}1}^{u_i+1}(v_d).$$ We are done with the first half of the proof. Namely, we are finished with the “combinatorial side” $`m_f`$ of Equation (9). Next we turn to the “algebraic side” $`E(u,v)`$ and show that it too is a sum of terms indexed by the sequences $`\underset{¯}{\mathrm{}}`$. We show that the term corresponding to a sequence $`\underset{¯}{\mathrm{}}`$ equals the right hand side of (12). This clearly suffices to complete the proof. By definition, $`E(u,v)`$ is the expression (3). The entries in the last column of the numerator vanish in rows $`i`$ for $`1ir1`$ because $`u_i+1u_{r1}+1v_d`$ and $`0=ϵ_{v_d}ϵ_{v_d}`$ occurs as a factor in $`\mu _n^{u_i+1}(v_d)`$. We would like to kill also the entries in rows $`r`$ through $`d1`$. To this end, subtract from row $`i`$, $`rid1`$, the multiple of row $`d`$ by $`\mu _{u_d}^{u_i+1}(v_d)`$. The entry in position $`(i,j)`$ then becomes $$\mu _n^{u_i+1}(v_j)\mu _{u_d}^{u_i+1}(v_d)\mu _n^{u_d+1}(v_j)=\mu _n^{u_d+1}(v_j)\left(\mu _{u_d}^{u_i+1}(v_j)\mu _{u_d}^{u_i+1}(v_d)\right).$$ In particular, all the entries in the last column except the one on row $`d`$ are $`0`$. The factor $`\mu _n^{u_d+1}(v_j)`$ being now common to all the entries in column $`j`$, let us take these factors out of every column. The resulting entries in the last column are all zero except the one on row $`d`$ which is $`1`$. The numerator therefore reduces to the determinant of the submatrix of the first $`d1`$ rows and columns. Let us take the factors $`(ϵ_{v_d}ϵ_{v_1})`$, …, $`(ϵ_{v_d}ϵ_{v_{d1}})`$ out of the denominator and distribute them thus: divide by $`(ϵ_{v_d}ϵ_{v_j})`$ every entry in column $`j`$ of the determinant in the numerator. After these manipulations, expression (3) looks like this: (13) $$\underset{j=1}{\overset{d}{}}\mu _n^{u_d+1}(v_j)\frac{det(A)}{ϵ(v_2,v_1)\left(ϵ(v_3,v_1)ϵ(v_3,v_2)\right)\mathrm{}\left(ϵ(v_{d1},v_1)\mathrm{}ϵ(v_{d1},v_{d2})\right)}$$ where $`A:=(A_{ij})`$ is the $`d1\times d1`$ matrix whose entry at position $`(i,j)`$ is (14) $$A_{ij}=\frac{\mu _{u_d}^{u_i+1}(v_j)\mu _{u_d}^{u_i+1}(v_d)}{ϵ_{v_d}ϵ_{v_j}}.$$ Now apply the following row operations to the matrix $`A`$: subtract row $`r+1`$ from row $`r`$, …, row $`d1`$ from row $`d2`$. To get a handle on the resulting matrix—let us call it $`B`$—we use the following equation which is proved readily by induction: for positive integers $`aeb`$, $`c`$, and $`f`$, we have $$\begin{array}{ccc}\hfill \frac{\mu _b^a(c)\mu _b^a(f)}{ϵ_fϵ_c}& =& \mu _b^{a+1}(c)+\mu _b^{a+2}(c)\mu _a^a(f)+\mathrm{}+\mu _b^e(c)\mu _{e2}^a(f)+\hfill \\ & & \frac{\mu _b^e(c)\mu _b^e(f)}{ϵ_fϵ_c}\hfill \\ & =& \left(\underset{a1<\mathrm{}e1}{}\mu _b^{\mathrm{}+1}(c)\mu _\mathrm{}1^a(f)\right)+\frac{\mu _b^e(c)\mu _b^e(f)}{ϵ_fϵ_c}\hfill \end{array}$$ Applying this to (14) for $`i`$ such that $`rid1`$ and $`e=u_{i+1}+1`$, we get $$\begin{array}{ccc}\hfill \frac{\mu _{u_d}^{u_i+1}(v_j)\mu _{u_d}^{u_i+1}(v_d)}{ϵ_{v_d}ϵ_{v_j}}& =& \left(\underset{u_i<\mathrm{}_{i+1}u_{i+1}}{}\mu _{u_d}^{\mathrm{}_{i+1}+1}(v_j)\mu _{\mathrm{}_{i+1}1}^{u_i+1}(v_d)\right)+\hfill \\ & & \frac{\mu _{u_d}^{u_{i+1}+1}(v_j)\mu _{u_d}^{u_{i+1}+1}(v_d)}{ϵ_{v_d}ϵ_{v_j}}\hfill \end{array}$$ where we have written $`\mathrm{}_{i+1}`$ rather than just $`\mathrm{}`$ for the running index in the sum. Note that the second term on the right is precisely the entry at position $`(i+1,j)`$ of $`A`$ for $`rid2`$ and vanishes for $`i=d1`$. Thus the entry at position $`(i,j)`$ of the matrix $`B`$ looks like this: $$B_{ij}=\{\begin{array}{cc}\frac{\mu _{u_d}^{u_i+1}(v_j)}{ϵ(v_d,v_j)}\hfill & \text{if }i<r\hfill \\ \underset{u_i<\mathrm{}_{i+1}u_{i+1}}{}\mu _{u_d}^{l_{i+1}+1}(v_j)\mu _{\mathrm{}_{i+1}1}^{u_i+1}(v_d)\hfill & \text{if }ir\text{.}\hfill \end{array}$$ By the multilinearity of the determinant, we see that $`det(B)`$ (which equals $`det(A)`$, since $`B`$ was obtained from $`A`$ by elementary row operations) equals the sum over $`\underset{¯}{\mathrm{}}=(\mathrm{}_{r+1},\mathrm{},\mathrm{}_d)`$ of (15) $$\underset{ri<d}{}\mu _{\mathrm{}_{i+1}1}^{u_i+1}(v_d)\left|\begin{array}{ccccc}\frac{\mu _{u_d}^{u_1+1}(v_1)}{ϵ(v_d,v_1)}& \mathrm{}& \frac{\mu _{u_d}^{u_1+1}(v_j)}{ϵ(v_d,v_j)}& \mathrm{}& \frac{\mu _{u_d}^{u_1+1}(v_{d1})}{ϵ(v_d,v_{d1})}\\ \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}\\ \frac{\mu _{u_d}^{u_{r1}+1}(v_1)}{ϵ(v_d,v_1)}& \mathrm{}& \frac{\mu _{u_d}^{u_{r1}+1}(v_j)}{ϵ(v_d,v_j)}& \mathrm{}& \frac{\mu _{u_d}^{u_{r1}+1}(v_{d1})}{ϵ(v_d,v_{d1})}\\ \mu _{u_d}^{l_{r+1}+1}(v_1)& \mathrm{}& \mu _{u_d}^{l_{r+1}+1}(v_j)& \mathrm{}& \mu _{u_d}^{l_{r+1}+1}(v_{d1})\\ \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}\\ \mu _{u_d}^{l_d+1}(v_1)& \mathrm{}& \mu _{u_d}^{l_d+1}(v_j)& \mathrm{}& \mu _{u_d}^{l_d+1}(v_{d1})\end{array}\right|.$$ Substitute this into equation (13). Multiplying the factor $`\mu _n^{u_d+1}(v_j)`$, for $`1jd1`$, in equation (13) into all the entries in column $`j`$ of the determinant in (15) yields the determinant in the numerator of $`E(u^{},v^{};v_d)`$ in (11). It should now be clear that $`E(u,v)`$ is the sum over $`\underset{¯}{\mathrm{}}`$ of the right side of (12), and the proof of the restriction formula (Theorem 2) is finally over. ## 6. Gröbner degeneration computes restriction The goal of this section is to state and prove Theorem 3 below, which was used in the proof in §5 of the restriction formula (Theorem 2). As pointed out to us by the referee, Theorem 3 is well known and can be deduced from results in the literature. The assumptions and notations of the previous sections are annulled now. Fix a torus $`T:=(^{})^m`$. Let $`Z`$ be a non-singular complex projective variety of dimension $`d`$ on which there is an algebraic action of $`T`$ with finitely many fixed points. Then, by Bialynicki-Birula , $`Z`$ admits an equivariant algebraic cell decomposition, and around each $`T`$-fixed point there is a $`T`$-stable open subset $`U`$ of $`Z`$ that is isomorphic to a $`T`$-module (the fixed point of course corresponds to $`0`$ in the module). Let $`Y`$ be a $`T`$-stable irreducible subvariety of $`Z`$, $`y`$ a $`T`$-fixed point on $`Y`$, and $`U^d`$ a $`T`$-stable open subset of $`Z`$ containing $`y`$ as above. Let $`X_1,\mathrm{},X_d`$ be coordinates on $`U`$ that are eigenvectors for $`T`$—since $`y`$ is an isolated fixed point, no coordinate has trivial character. Choose some term order on the set of monomials in the coordinates, and suppose that $`J`$ is the initial ideal, with respect to this order, of the ideal of functions on $`U`$ vanishing on $`YU`$. Since $`J`$ is a monomial ideal, it has a primary decomposition consisting of monomial ideals. Let $`_{i=1}^pJ_i`$ be the intersection of the minimal primary components (we are throwing away the embedded components). The radical $`𝔭_i`$ of $`J_i`$ is of the form $`(X_1^{a_{1i}},\mathrm{},X_d^{a_{di}})`$ where each $`a_{ji}`$ is either $`0`$ or $`1`$ (exactly $`dimY`$ of the $`a_{ji}`$ equal $`0`$ for each $`i`$). The scheme $`M`$ defined by $`_{i=1}^pJ_i`$ is the union $`_{i=1}^pM_i`$ of the schemes $`M_i`$ defined by $`J_i`$. Let $`\mathrm{}_i`$ be the length of $`R_{𝔭_i}/JR_{𝔭_i}=R_{𝔭_i}/J_iR_{𝔭_i}`$ where $`R:=[X_1,\mathrm{},X_d]`$. ###### Theorem 3. With hypothesis and notation as above, the restriction to the fixed point $`y`$ of the equivariant cohomology class $`[Y]`$ of the subvariety $`Y`$ in the equivariant integral cohomology ring of $`Z`$ is given by (16) $$[Y]|_y=\underset{i=1}{\overset{p}{}}\mathrm{}_i\underset{j=1}{\overset{d}{}}\chi _j^{a_{ji}}$$ where $`\chi _1`$, …, $`\chi _d`$ are respectively the characters of $`X_1`$,…, $`X_d`$. Proof : We thank the referee for indicating how the theorem can be deduced as follows from known results. The restriction of the equivariant Chow class of $`Y`$ to the fixed point $`y`$ factors through the restriction to the open set $`U`$. The fact that the class of $`YU`$ and of $`M`$ in the equivariant Chow ring of $`U`$ are the same and equal to the right hand side of (16) can be found in any number of references under the heading of “equivariant Chow”, or “multidegree”, or “equivariant Hilbert polynomials”, or “equivariant multiplicity.” See, for example, \[11, Notes to Chapter 8\]; the fact that the “multidegrees” in the above reference are equivariant cohomology classes is asserted in Proposition 1.19 of , where multidegrees are identified as being equivariant Chow classes. $`\mathrm{}`$ ## 7. Towards an equivariant Pieri formula Recall, from \[3, Eq. (9), p.146\] for example, that the classical Pieri formula gives a beautiful expression, as an integral linear combination of general Schubert classes, for the product of a special Schubert class with a general Schubert class in the ordinary cohomology ring of the Grassmannian. It seems like there ought to be a similarly beautiful closed-form equivariant version that specializes to the ordinary one. Unfortunately, this we do not yet have. All we want to do in this section is record an observation (see Proposition below) that is a formal consequence of the following basic and well-known facts: the injection of Equation (1); the restriction to a $`T`$-fixed point of an (equivariant) Schubert class vanishes if the fixed point is not contained in the Schubert variety; and the degree of a Schubert class equals the codimension of the Schubert variety. Being a formal consequence, the observation holds for any generalized flag variety, not just the Grassmannian. The point to note is that the right hand side of (17) is in terms of restrictions, which, thanks to Theorem 2, we know how to compute in the case of Grassmannians. Let $`G`$ be a complex semisimple algebraic group and $`Q`$ a parabolic subgroup. Let $`T`$ be a maximal torus and $`B`$ a Borel subgroup of $`G`$ such that $`TBQ`$. Let $`W`$ denote the Weyl group of $`G`$ with respect to $`T`$, and $`W_Q`$ the Weyl group of (the Levi part of) $`Q`$ with respect to $`T`$. The Schubert varieties in $`G/Q`$ are by definition the $`B`$-orbit closures for the action of $`B`$ on $`G/Q`$ by left multiplication. They are naturally indexed by $`W/W_Q`$. We use $`u`$, $`v`$, $`w`$, … to denote elements of $`W/W_Q`$; $`X(u)`$, $`X(v)`$, $`X(w)`$, … denote the corresponding Schubert varieties; $`[X(u)]`$, $`[X(v)]`$, $`[X(w)]`$, … denote the corresponding equivariant Schubert classes. The partial order on Schubert varieties by inclusion induces a partial order, denoted $``$, on $`W/W_Q`$. Let $`c_{uv}^w`$ be the structure constants of the multiplication of the equivariant integral cohomology ring of $`G/Q`$ with the respect to the basis of Schubert classes: $$[X(u)][X(v)]=\underset{w}{}c_{uv}^w[X(w)].$$ The proof of the following proposition is a straightforward induction argument and so we omit it. ###### Proposition 4. 1. $`c_{uv}^w=0`$ unless $`wu`$, $`wv`$, and the codimensions are such that $`\text{codim }X(u)+\text{codim }X(v)\text{codim }X(w)`$. 2. If $`wu`$ and $`wv`$, then (17) $$c_{uv}^w=\underset{w=y_0<\mathrm{}<y_k}{}(1)^k\frac{[X(u)]|_{y_k}[X(v)]|_{y_k}}{[X(y_k)]|_{y_k}}\frac{[X(y_k)]|_{y_{k1}}}{[X(y_{k1})]|_{y_{k1}}}\mathrm{}\frac{[X(y_1)]|_{y_0}}{[X(y_0)]|_{y_0}}$$ where the sum is over all chains $`w=y_0<\mathrm{}<y_k`$ with $`y_ku`$ and $`y_kv`$; $`[X(u)]|_{y_k}`$ denotes the restriction of the Schubert class $`[X(u)]`$ to the $`T`$-fixed point indexed by $`y_k`$. $`\mathrm{}`$
warning/0506/gr-qc0506023.html
ar5iv
text
# Untitled Document SCALAR FIELD ENTROPY IN BRANE WORLD BLACK HOLES K.K. Nandi<sup>1,2,a</sup>, Y.-Z. Zhang<sup>2,b</sup>, A. Bhadra<sup>4,c</sup> and P. Mitra<sup>5,d</sup> <sup>1</sup>Department of Mathematics, University of North Bengal, Darjeeling (W.B.) 734 430, India <sup>2</sup>Institute of Theoretical Physics, Chinese Academy of Sciences, P.O.Box 2735, Beijing 100080, China <sup>4</sup>IRC, University of North Bengal, Darjeeling (W.B.) 734 430, India <sup>5</sup>Theoretical Nuclear Physics Division, Saha Institute of Nuclear Physics, 1/AF Bidhan Nagar, Salt Lake, Calcutta 700 064, India ————————————————————– <sup>a</sup>E-mail address: kamalnandi1952@yahoo.co.in <sup>b</sup>E-mail address: yzhang@itp.ac.cn $`\text{}^c`$E-mail address: aru\_bhadra@yahoo.com <sup>d</sup>E-mail address: parthasarathi.mitra@saha.ac.in Abstract A semiclassical calculation of entropy of a scalar field in the background of a class of brane world black holes (BWBH) is carried out in the presence of a brick wall cutoff when the 5D-bulk induced “tidal charge” has generic or extreme values. ———————————— The entropy $`S`$ of a standard nonextremal black hole obeys the well known Bekenstein-Hawking area law $`S=\frac{A}{4G}`$, where $`A`$ is the area of the horizon. This result can be rederived by the Lagrangian path integral formalism or by other means . However, the entropy due to quantum fields in the black hole background introduces divergences which are interpreted as renormalizations of the gravitational coupling constant $`G`$. In the extreme case defined by two coinciding surfaces, the situation is different - the area law does not hold ($`S=0`$). It has been argued that extreme and nonextreme BHs should be regarded as qualitatively different objects due to discontinuity in the Euclidean topology . In this context, one would recall that dilatonic black holes have zero area at extremality and hence $`S=0`$. However, $`S`$ does not remain zero when quantum contributions are taken into account: The linear divergence goes away but a nonleading logarithmic divergence persists . Given this scenario and given the current widespread interest in the brane theory, it would be important to assess similar quantum contributions to entropy in the background of BWBHs. The purpose of the present note is to do this. The motivation is that, generically, the BWBHs are different from ordinary BHs as they embody a synthesis of wormhole and blackhole features. For instance, the effective stress energy tensor violates some of the energy conditions. As argued in Refs., this feature is not unexpected as the tensor contains imprints of the free gravitational field in the bulk which contributes negative energy. Essentially, the bulk contribution implies a correction to the Schwarzschild solution but its horizon structure remains undisturbed. The brane theory we have in mind is described by the RS2 framework, that is, a single brane in a $`Z_2`$-symmetric 5-D asymptotically anti-de Sitter bulk in which only gravity propagates while all other fields are confined to the brane . The 5-D Weyl tensor projected onto the brane is represented by a trace-free tensor $`E_\mu ^\nu `$ appearing in the Shiromizu-Maeda-Sasaki equations . The trace of these equations, viz., Ricci scalar $`R=0`$, are solved to derive the BWBHs. Henceforth, we take units $`G=c=\mathrm{}=k=1`$, and Greek indices run from $`0`$ to $`3`$. The general class of 4-D solutions is given by $$ds^2=g_{\mu \nu }dx^\mu dx^\nu =A(r)dt^2+\frac{dr^2}{B(r)}+r^2d\mathrm{\Omega }^2$$ (1) in which $`d\mathrm{\Omega }^2`$ is the metric on the unit $`2`$-sphere, $`A(r)`$ and $`B(r)`$ are two well behaved positive functions for $`r>r_h`$ and have a simple zero at $`r=r_h`$ defining the horizon. The singularities, if any, of the BH solutions when propagated off the brane into the 5-D bulk may make the AdS horizon singular (“black cigar” ). However, several classes of nonsingular, static, spherically symmetric BWBH solutions have now been proposed almost simultaneously \[13-15\] and some quantum properties have also been investigated . Under certain assumptions on the behavior of the metric functions, Bronnikov, Melnikov and Dehnen have shown that the solution (1) can have $`R\times S^2`$ topology of spatial sections. Assuming asymptotic flatness at large $`r`$, the global causal structure of such solutions coincides with that of Kerr-Newman nonextremal solutions. We shall specifically consider a class of solutions given by : $$A(r)=1\frac{2M}{r};B(r)=\frac{(1\frac{2M}{r})(1\frac{r_0}{r})}{1\frac{3M}{2r}}$$ (2) where $`M`$ and $`r_0(0)`$ are two adjustable parameters, one is related to the mass and the other may be interpreted as an induced tidal charge - a Weyl tensor projection from the bulk into the brane. The ranges $`r_0>2M,r_0=2M`$ and $`r_0<2M`$ correspond respectively to traversable wormholes, extremal and nonextremal regular BWBHs. The areas of both categories of black holes are nonzero. For $`r_0=3M/2`$, one recovers the Schwarzschild solution. Let us consider the solution (2) and calculate the background values of $`S`$ and $`\kappa `$ following : $$S=\frac{1}{8\pi }_{Hor}[K]=\frac{A(r_h)}{4},$$ (3) where $`[K]=KK_0`$ in which $`K`$ is the extrinsic curvature and $`K_0`$ is obtained by substituting into $`K`$ the flat space metric. The surface gravity $`\kappa `$ at the horizon $`r_h=2M`$ is given by $$\frac{\kappa }{2\pi }=\frac{1}{\beta }=\frac{1}{4\pi }\frac{_rg_{00}}{\sqrt{g_{rr}g_{tt}}}=\frac{1}{4\pi M}\left(1\frac{r_0}{2M}\right)^{\frac{1}{2}}.$$ (4) We shall now adopt the statistical mechanics approach via the partition function for the quantization of a scalar field field $`\varphi `$ in the fixed background given by the solution (1). The dominant contribution to the partition function $`Z`$ will be assumed to come from the classical solutions of the Euclidean Lagrangian action $`S_1[g_{cl}]`$ leading to the area law in general. Neglecting the quantum fluctuations of the metric and other bulk induced quantities, that is, freezing them into their background values, the partition function can be represented as $$Z=e^{S_1[g_{cl}]}[d\varphi ]e^{S_2[g_{cl},\varphi ]}.$$ (5) In order to find the contribution of $`\varphi `$ to the partition function $`Z`$ through $`S_2`$, we further employ the “brick wall” boundary condition of ’t Hooft . That means, we take $`\varphi (r)=0`$ at $`r=2M+ϵ`$ where the small quantity $`ϵ`$ signifies an ultraviolet cut-off. There is also an infrared cut-off, that is, $`\varphi (r)=0`$ at $`r=L`$ where $`L>>2M`$. The Klein-Gordon equation for this scalar field in the background of the spacetime $`g_{\mu \nu }`$ reads $$\frac{1}{\sqrt{g}}_\mu \left(\sqrt{g}g^{\mu \nu }_\nu \varphi \right)m^2\varphi =0.$$ (6) An ansatz of the form $`\varphi =e^{iEt}f_{El}Y_{lm}`$ yields the radial equation for the metric (1): $$A^1E^2f_{El}+\left(\frac{B}{A}\right)^{\frac{1}{2}}\frac{1}{r^2}\frac{}{r}\left[\left(AB\right)^{\frac{1}{2}}r^2\frac{f_{El}}{r}\right]\left[l(l+1)+m^2\right]f_{El}=0.$$ (7) Taking further $`f_{El}e^{i{\scriptscriptstyle 𝑑rp}}`$, and keeping the real parts, a radial wave number can be found as $$p^2=B^1\left[A^1E^2l(l+1)m^2\right].$$ (8) Restricting $`p`$ to the real values, we introduce the semiclassical quantization condition $$\pi n_r=_{2M+ϵ}^Lp(r,l,E;m)𝑑r$$ (9) where $`n_r`$ must be a nonnegative positive integer. Then, with $`\beta =1/T=2\pi /\kappa `$, where $`T`$ is the temperature, $`\kappa `$ is the surface gravity, the Helmholtz free energy $`F`$, appearing in $`Z=e^{\beta F}`$, is given by $`\beta F`$ $`=`$ $`{\displaystyle \underset{n_r,l,m}{}}\mathrm{ln}(1e^{\beta E}){\displaystyle 𝑑l(2l+1)\mathrm{ln}(1e^{\beta E})𝑑n_r}`$ (10) $`=`$ $`{\displaystyle 𝑑l(2l+1)d(\beta E)\left(e^{\beta E}1\right)^1n_r}`$ $`=`$ $`{\displaystyle \frac{2\beta }{3\pi }}{\displaystyle _{2M+ϵ}^L}(A)^{3/2}(B)^{1/2}r^2𝑑r{\displaystyle 𝑑E\left(e^{\beta E}1\right)^1\left[E^2Am^2\right]^{\frac{3}{2}}},`$ where the $`l`$ integration has been explicitly carried out. Now, as is customary , we set the lower limit in the $`E`$ integral to near zero and evaluate it approximately ignoring the contributions coming from the $`m^2\beta ^2`$ term. Then we find the following result for $`M0`$ from Eq.(8): There is a linearly divergent contribution for $`ϵ0`$ given by $$F_{corr}\frac{4\pi ^3}{180ϵ}\left(\frac{2M}{\beta }\right)^4\left(1\frac{r_0}{2M}\right)^{1/2}.$$ (11) The entropy $`S`$ can be expressed in terms of the free energy as $$S=\beta ^2\frac{F}{\beta }$$ (12) and the contribution from the linearly divergent term is given by $$S_{corr}=\frac{16\pi ^3}{180}\left(\frac{2M}{\beta }\right)^3\frac{\left(1\frac{r_0}{2M}\right)^{1/2}}{2Mϵ}(2M)^2.$$ (13) At the horizon, on putting the value of $`\beta `$ from (4), this reduces to $$S_{corr}=\frac{1}{90}\frac{\left(1\frac{r_0}{2M}\right)}{2Mϵ}(2M)^2.$$ (14) Thus the nonextremal ($`r_02M`$) behavior of the cut-off dependence ($`ϵ0`$) is similar to that in other black holes. It is important to introduce the proper thickness $`\stackrel{~}{ϵ}`$ of the black hole at this point: $$\stackrel{~}{ϵ}=_{2M}^{2M+ϵ}\frac{dr\sqrt{1\frac{3M}{2r}}}{\sqrt{1\frac{2M}{r}}\sqrt{1\frac{r_0}{r}}}=\frac{\sqrt{2Mϵ}}{\sqrt{1\frac{r_0}{2M}}}.$$ (15) The entropy $`S_{corr}`$ can be rewritten as $$S_{corr}=\frac{1}{90}\frac{1}{\stackrel{~}{ϵ}^2}(2M)^2.$$ (16) Thus, if we keep $`\stackrel{~}{ϵ}`$ fixed, we find that, in the extremal limit $`r_02M`$, $`S_{corr}`$ remains finite. This is so in spite of the vanishing of Hawking temperature in this limit (Actually the factor $`(1\frac{r_0}{2M})`$ cancels out). The reason for this is that the divergence in the $`r`$ integral becomes stronger and compensates the temperature effect. The physical difference between extreme and near extreme BHs have been discussed at length in Ref.. A direct calculation may be carried out for the extremal case. In this situation, the expression for the temperature given above vanishes, but the Euclidean black hole has no conical singularity and the time can be given any periodicity, so that the temperature is arbitrary. With this understanding, one can write the contribution from the field to the entropy as $$S_{corr}^{ex}=\frac{8}{135}\pi ^3\left(\frac{2M}{\beta }\right)^3\left(\frac{2M}{ϵ}\right)^{3/2}=\frac{8}{135}\pi ^3\left(\frac{2M}{\beta }\right)^3\mathrm{exp}(\frac{3}{2}\mathrm{\Lambda }),$$ (17) where $`\mathrm{\Lambda }`$ is a large cutoff representing the proper distance from the horizon (at negative infinity in $`r`$) to the brick wall . This expression shows much stronger exponential divergence, as compared to the quadratic divergence shown in the extreme limit of the non-extremal case. The above results for the BWBH may be compared with the thermodynamics of the Reissner-Nordström solution. In both cases, the period $`\beta `$ in Euclidean time is infinite in the extremal limit and arbitrary at exact extremality, giving zero entropy for the purely gravitational theory (see Refs. for a different approach to quantization). The nonextremal contributions of the matter field contain mild divergences, but the extremal limit does not produce anything stronger: $`S_{corr}`$ goes to a finite limit. The extremal black holes, if treated directly, show a stronger divergence. In the dilatonic solution, the area law holds and in the extremal limit or extremal case the leading contribution to $`S_{corr}`$ disappears since $`A=0`$, but leaves a logarithmic divergence . We can extend the approach in this paper to another type of BWBH for which the metric functions are \[7,8,13-15\] $$A(r)=B(r)=1\frac{2M}{r}\frac{Q^2}{r^2},$$ (18) in which $`M`$ is the mass and $`Q`$ is the tidal charge. Due to the wrong sign before $`Q^2>0`$, the stress energy violates the energy conditions. It should be remarked that, historically, this kind of solution was heuristically conceived by Einstein and Rosen in their attempt to construct singularity free particle models (“Einstein-Rosen bridge”) . However, the brane world physics of today provides a physical justification for this sign reveresed solution via tidal effects albeit the spacetime shares some properties of a wormhole (two sheeted topology, for instance). Nevertheless, the solution is treated as a BWBH and is recently applied in the context of gravitational lensing . The BH horizons appear at $`r_h^\pm =M\pm \sqrt{M^2+Q^2}`$ and the temperature of the outer horizon $`r_h^+`$ is given by $$T=\frac{\sqrt{M^2+Q^2}}{2\pi \left[M+\sqrt{M^2+Q^2}\right]^2}$$ (19) and the area law is satisfied. But, $`T=0`$ is possible only if $`M=0`$ and $`Q=0`$, that means, a flat space. Thus, nontrivial extreme solution like the ones dealt with above simply does not exist here! Calculations show that there occur both linear and logarithmic divergences including in the case $`M=0`$ (pure tidal charge solution). We withhold the details except only mentioning that these divergences persist also in all other types of BWBHs proposed in Ref.. In this sense, the first example occupies a very special place indded. To summarize, we have brought out the quantum behavior of a scalar field in the background of BWBHs. It is found that nonextremal BWBHs produce ultraviolet ($`ϵ0`$) divergences in entropy, as do other ordinary BHs. The nontrivial extremal BWBH produces stronger (exponential) divergences like in the RN extreme case. This is perhaps understandable noting that both BWBH and RNBH satisfy the same field equation $`R=0`$, where $`R`$ is the Ricci scalar. However, the extreme counterpart does not exist in all classes of BWBH with tidal charge. These results underline the distinctive features of BWBH thermodymamics. ACKNOWLEDGMENT Part of the work was carried out while one of the authors (KKN) was visiting Ufa, Russia. It is a pleasure to thank Guzel Kutdusova for several useful assistance. This project was in part supported by NNSFC under Grant No. 90403032 and also by National Basic Research Program of China under Grant No. 2003CB716300. REFERENCES G.W. Gibbons and S.W. Hawking, Phys. Rev. D 15, 2752 (1977). L. Susskind and J. Uglum, Phys. Rev. D 50, 2700 (1994). R. Kallosh, T. Ortín, and A. Peet, Phys. Rev. D 47, 5400 (1993). For a recent discussion in a general context, see: T. Padmanabhan, Mod. Phys. Lett. A 19, 2637 (2004); Class. Quant. Grav. 21, 4485 (2004). S.W. Hawking, G.T. Horowitz, and S.F. Ross, Phys. Rev. D 51, 4302 (1995). G.W. Gibbons and K. Maeda, Nucl. Phys. B 298, 741 (1988); D. Garfinkle, G. Horowitz, and A. Strominger, Phys. Rev. D 45, 3888 (1992). A. Ghosh and P. Mitra, Phys. Rev. Lett. 19, 2521 (1994). N. Dadhich, Phys. Lett. B 492, 357 (2000); N. Dadhich, R. Maartens, P. Papadopoulos, and V. Rezania, Phys. Lett. B 487, 1 (2000). S. Shankarnarayanan and N. Dadhich, Int. J. Mod. Phys. D 13, 1095 (2004). L. Randall and R. Sundrum, Phys. Rev. Lett. 83, 4690 (1999). T. Shiromizu, K. Maeda, and M. Sasaki, Phys. Rev. D 62, 024012 (2000). K.A. Bronnikov, V.N. Melnikov and H. Dehnen, Phys. Rev. D 68, 024025 (2003). A. Chamblin, S.W. Hawking, and H.S. Reall, Phys. Rev. D 61, 065007 (2000). R. Casadio, A. Fabbri, and L. Mazzacurati, Phys. Rev. D 65, 084040 (2002). See also Refs.. N. Dadhich, S. Kar, S. Mukherjee, and M. Visser, Phys. Rev. D 65, 064004 (2002). Exterior metric similar to Eq.(19), but applicable for a star in the brane, has been derived by: C. Germani and R. Maartens, Phys. Rev. D. 64, 124010 (2001). They also derived the interior metric. R. Casadio and B. Harms, Phys. Lett. B 487, 209 (2000); R. Casadio, Ann. Phys. (NY) 307, 195 (2003); Phys. Rev. D 69, 084025 (2004). G. ’t Hooft, Nucl. Phys. B 256, 727 (1985). A. Ghosh and P. Mitra, Phys. Lett. B 357, 295 (1995). A. Ghosh and P. Mitra, Phys. Rev. Lett. 78, 1858 (1997). For a grand canonical ensemble approach involving charged black holes, see: H.W. Braden, J.D. Brown, B.F. Whiting, and J.W. York, Phys. Rev. D 42, 3376 (1990). J. Preskill, P. Schwarz, A. Shapere, S. Trivedi, and F. Wilczek, Mod. Phys. Lett. A 6, 2353 (1991); C. Holzhey and F. Wilczek, Nucl. Phys. B 380, 447 (1992); R. Kallosh, A. Linde, T. Ortín, and A. van Proeyen, Phys. Rev. D 46, 5278 (1992). A. Einstein and N. Rosen, Phys. Rev. 48, 73 (1935). R. Whisker, Phys. Rev. D 71, 064004 (2005).
warning/0506/math0506626.html
ar5iv
text
# 1 Introduction ## 1 Introduction ### 1.1 Motivation The simplex algorithm is widely used to solve linear programs. It works well in practice, though often one cannot prove that it will. A parameter in the algorithm is the rule that selects one move among possible moves (“pivots”) that decrease the objective function. Deterministic pivot rules are known to be possibly very far from optimal. For example, consider problems with the number $`m`$ of constraints of the same order as the dimension, $`n`$. For virtually every deterministic pivot rule there is a problem for which the algorithm will take exponential time, although it is conjectured that there exists a descent path whose length $`O(n)`$. Variants of the original argument by Klee and Minty \[KM72\] are cited in \[GHZ98, page 2\]. Randomized pivot rules appear to do better. According to \[GHZ98\], several of the most popular randomized pivot rules appear to have polynomial – even quadratic – running time. Rigorous and general results on these, however, have been hard to come by. When one restricts to a narrow class of test problems, it becomes possible to obtain some rigorous results. Gärtner, Hank and Ziegler \[GHZ98\] consider three randomized pivot rules. Relevant to the present paper are their results on the random edge rule, in which the next move is chosen uniformly among edges leading to decrease the objective function. They analyze the performance of this rule on a class of linear programs, the feasible polyhedra for which are called Klee-Minty cubes, after \[KM72\]. Such cubes are good benchmarks because they cause some pivot rules to pass through a positive fraction of the vertices. They prove that the expected run time is quadratic, up to a possible log factor in the lower bound: ###### Theorem 1 (GHZ) The expected number, $`E_n`$ of steps taken by the random edge rule, started at a random vertex of a Klee-Minty cube, is bounded by $$\frac{n^2}{4H_{n+1}1}E_n\left(\genfrac{}{}{0pt}{}{n+1}{2}\right).$$ Here, $`H_n=_{j=1}^n\frac{1}{j}\mathrm{log}n`$ is the $`n^{th}`$ harmonic number. Their lower bound rules out the possibility that $`E_nn\mathrm{polylog}(n)`$ which was twice conjectured by previous researchers \[PS82, page 29\]\[Kel81\]. They guess that the upper bound is the correct order of magnitude, and state an improvement in the upper bound from $`(1/2)n^2`$ to $`0.27\mathrm{}n^2`$, whose proof is omitted. The method of analysis in \[GHZ98\] is via a combinatorial model due to \[AC78\] which describes the progress of the algorithm as a random walk on an acyclic directed graph. In their model, vertices are bijectively mapped to sequences of 0’s and 1’s of length $`n`$, and each move consists of flipping a 1 (chosen uniformly at random) to a 0, and simultaneously flipping all bits to the right of the chosen bit. It was in this form that the problem came to our attention. Indeed, the remainder of the paper is framed in terms of a variant of this model, which we find to be an intrinsically interesting model. Our main result, Theorem 2, closes the gap left open in \[GHZ98\], proving that the upper bound is sharp to within a constant factor and obtaining upper and lower bounds differing by a factor of less than 2. We have moved the model to continuous time and made $`n`$ infinite, since from our view as probabilists this is the most natural way to frame such a model. Nevertheless, our results apply to the setting of \[GHZ98\] as well (Corollary 3). Although the model seems simple, we remark that we were unable to prove many things about the model, including whether certain limits exist. ### 1.2 Statement of the model The one-dimensional integer lattice is decorated with 0’s and 1’s, arbitrarily except that there must be some point to the left of which lie only 0’s. Each site has a clock that goes off at times distributed as independent mean-one exponentials. When a clock rings, if there is a 0 there nothing happens, but if there is 1 there, then it and all (infinitely many) of the sites to the right flip as well – 0’s become 1’s and 1’s become 0’s. Later arguments will use random variables involved in the construction of this continuous time Markov chain, so we give a formal construction as follows. Let $`𝒮`$ be the subset of $`\{0,1\}^{}`$ consisting of sequences of 0’s and 1’s that have a leftmost 1 (equivalently, have finitely many 1’s to the left of the origin). Let $`\{N(j,t):t0\}_j`$ be a collection of IID Poisson counting processes, that is, step functions increasing by 1 at random times, which we denote $`\xi _{j,i}`$, having independent increments distributed as exponentials of mean 1. As usual, the filtration is defined by letting $`_t`$ denote the $`\sigma `$-field generated by $`\{N(j,s):st\}`$. For any $`\omega 𝒮`$, we define a Markov chain starting from $`\omega `$ as follows. If $`\omega `$ is the zero string, the chain remains at $`\omega `$. Otherwise, let $`i`$ be the position of the leftmost 1 in $`\omega `$. First, fix $`j>i`$ and define a Markov chain on configurations on the index set $`\{i,i+1,\mathrm{},j\}`$ in the finitary model: each site $`s`$ attempts to flip at times $`\xi _{s,r}`$ for all $`r0`$. The flip is successful if and only if there is a 1 in position $`s`$ at time $`\xi _{s,r}^{}`$, in which case sites $`s+1,\mathrm{}j`$ flip as well. This is a well defined process because for each $`T`$ there are only finitely many jump times $`\xi _{s,r}`$ with $`isj`$ and $`rT`$. It is clear that for $`j^{}>j`$ the Markov chain thus defined on configurations on $`[i,j^{}]`$ projects down to the Markov chain on configurations on $`[i,j]`$ by ignoring the bits to the right of $`j`$. By Kolmogorov’s Extension Theorem there is an inverse limit as $`j\mathrm{}`$. This is a Markov chain, which we denote $`\{\omega _t\}`$, taking values in trajectories $`\omega :[0,\mathrm{})𝒮`$, and having jumps at a countable set of times $`\xi _{s,r}`$, at which a 1 flips to a 0 and all bits to its right flip as well. We are interested in the speed at which the leftmost 1 drifts to the right. Because this chain converges weakly to the zero state, it is convenient to renormalize by shifting to the leftmost 1. Consequently, we define the space $`\mathrm{\Xi }`$ to be the space of those sequences of 0’s and 1’s indexed by the nonnegative integers that begin with a 1. We define two functionals zeros and ones on $`\mathrm{\Xi }`$ by letting $`\text{ones}(𝐱)1`$ be the number of leading 1’s: $$\text{ones}(𝐱):=inf\{j1:𝐱(j)=0\}$$ and letting $`\text{zeros}(𝐱)0`$ be the number of successive 0’s after the first 1: $$\text{zeros}(𝐱):=1+inf\{j1:𝐱(j)=1\}.$$ Let $`\text{ones}_j`$ denote the set $`\{𝐱:\text{ones}(𝐱)=j\}`$ which form a partition of $`\mathrm{\Xi }`$, and let $`\{\text{zeros}_j\}`$ denote the analogous partition with respect to the values of zeros. We now define a Markov chain on the space $`\mathrm{\Xi }`$ whose law starting from $`𝐱\mathrm{\Xi }`$ is denoted $`𝐐_𝐱`$. Pick $`\omega 𝒮`$ such that the leftmost 1 of $`\omega `$ is in some position $`i`$ and $`\omega (i+j)=𝐱(j)`$ for all $`j0`$. We construct the Markov chain $`\{X_t\}`$ on $`\mathrm{\Xi }`$ as a function of $`\{M_t\}`$ as follows. First, define $`\sigma _0`$ to be 0, and $`i_0`$ to be the position of the leading 1 in $`\omega `$. Now recursively we let $`\sigma _n`$ be the first time after $`\sigma _{n1}`$ for which $`N(i_{n1},)`$ increases. For $`\sigma _{n1}t<\sigma _n`$ we let $`X_t`$ be $`M_t`$ shifted so that $`i_{n1}`$ is at the origin (and ignoring negative indices). We let $`\text{jump}_n`$ denote $`\text{ones}(\sigma _n^{})`$ and we let $`i_n=i_{n1}+\text{jump}_n`$. Sometimes, it is convenient to look at the chain sampled at times $`\sigma _n`$. Thus we let $$Y_n:=X_{\sigma _n}$$ which is now a discrete-time Markov chain. We let $`_𝐱`$ denote the law of this chain starting from $`𝐱`$. Since the conditional distribution of $`\sigma _n\sigma _{n1}`$ given $`_{\sigma _{n1}}`$ is exponential of mean 1, it follows that $`\sigma _n/n1`$. The distance the leading 1 has moved to the right by time $`t`$ is the sum $`_{n:\sigma _nt}\text{jump}_n`$, and therefore the average speed $`\text{spd}(n)`$ up to time $`\sigma _n`$ is the random quantity $$\text{spd}(n):=\sigma _n^1\underset{j=1}{\overset{n}{}}\text{jump}_jn^1\underset{j=1}{\overset{n}{}}\text{jump}_j.$$ (1) It is not a priori clear, nor in fact can we prove that $`\text{spd}(n)`$ has a limit as $`n\mathrm{}`$. Consequently we define inf-spd $`:=`$ $`\mathrm{liminf}_n\text{spd}(n);`$ sup-spd $`:=`$ $`\mathrm{limsup}_n\text{spd}(n).`$ > Problem 1: Show that the limiting speed exists. The Markov chain $`\{Y_n\}`$ is a time-homogeneous process on a compact space, so taking cesaro averages of the marginals, we see these must have at least one weak limit. Any weak limit is a stationary distribution. > Problem 2: Show that there is a unique stationary distribution for the chain $`\{Y_n\}`$. This would imply a positive solution to Problem 1. In particular, it would imply that $$n^1\underset{j=1}{\overset{n}{}}\text{jump}_j\text{jump}_1𝑑\pi $$ where $`\pi `$ is the unique stationary measure; hence the speed would not only exist but would be almost surely constant independent of the starting state. Although we do not have a solution to Problem 2, we believe something stronger may hold. > Problem 3a: Let $`T^j\pi `$ denote the composition of the measure $`\pi `$ with a translation by $`j`$ bits, e.g., if $`A`$ is the event that there is a 1 in position $`r`$, then $`T^j\pi (A)=\pi (A^j)`$, where $`A_j`$ is the event that there is a 1 in position $`r+j`$. Prove that $`T^j\pi M`$ where $`\pi `$ is the stationary measure and $`M`$ is IID fair coin flipping. > > Problem 3b: Prove or disprove that the unique stationary measure $`\pi `$ is equivalent (mutually absolutely continuous) to $`M`$. To see why we believe (3a) to be true, consider a window of positions of any size $`k`$, located far to the right of the initial 1. Between each time a clock rings in this window there are many times a 1 turns to a 0 to the left of the window. Each time this happens, the bits in the window all flip. Projecting configuration space onto what is visible in the window, and again onto a space of half the size by identifying each configuration with its complement, it seems reasonable to approximate the projection by a Markov chain. Specifically, from a state $`\{x,x^c\}`$, for each position $`i[1,\mathrm{},k]`$, exactly one of $`x`$ or $`x^c`$ will be able to flip the bit in position $`i`$, so one may imagine the pair $`\{x,x^c\}`$ as flipping at rate $`1/2`$ in every position. This chain has the uniform distribution as the unique stationary distribution. ## 2 Statement of main result and lemmas In this section we state the results that we do know how to prove, namely bounds on the lim inf and lim sup speeds. The following theorem is to be interpreted as referring to the lim inf and lim sup speeds (until we have proved the speed exists). ###### Theorem 2 The speed of the drift of the leftmost 1 satisfies $$1.646<\text{inf-spd}<\text{sup-spd}<2.92.$$ Relating back to the performance of the random edge rule on Klee-Minty cubes, we have: ###### Corollary 3 For sufficiently large $`n`$, starting from a uniform random vertex of the Klee-Minty cube, $$0.086n^2E_n0.152n^2.$$ Proof: Gärtner, Henk and Ziegler consider another way of counting steps, where instead of choosing an edge at random among all those decreasing the objective function, they choose an edge at random from among all edges, but suppress the move if the edge increases the objective function. For a vector $`𝐱`$ of 0’s and 1’s of length $`n`$, let $`L^{}(𝐱)=L^{}(𝐱,n)`$ denote the expectation of the number $`N^{(r)}`$ of moves starting from $`𝐱^r`$, including the suppressed moves, before the minimum is reached. They prove the following identity \[GHZ98, Lemma 4\]. $$E_n=\frac{1}{2n}\underset{r=1}{\overset{n}{}}L^{}(𝐱^r,n)$$ (2) where $`𝐱^r`$ is the vector of length $`r`$ consisting of all 1’s. Including suppressed moves in the count corresponds in our infinite, continuous-time model to counting the number of clock events (only among the first $`r`$ vertices). Let $`T^{(r)}`$ be the time it takes starting from $`𝐱^r`$ to reach the minimum. By the strong law of large numbers, $`N^{(r)}/T^{(r)}r`$ as $`r\mathrm{}`$. Also by the strong law, if the liminf and limsup speed are known to be in the interval $`(a,b)`$, then $$\frac{r}{b}<T^{(r)}<\frac{r}{a}$$ for sufficiently large $`r`$. Hence, for sufficiently large $`r`$, $$\frac{nr}{b}<N^{(r)}<\frac{nr}{a}$$ and plugging into (2) and summing from $`r=1`$ to $`n`$ gives $$\frac{n^2}{4b}<E_n<\frac{n^2}{4a}$$ for sufficiently large $`n`$. Plugging in $`a=2.92`$ and $`b=1.646`$ proves the corollary. $`\mathrm{}`$ The lower bound of Theorem 2 is proved in Section 4. The lower bound may in principle be improved so as to be arbitrarily near the actual speed. For the upper bound, we state some lemmas. Let $$H_k:=\underset{j=1}{\overset{k}{}}\frac{1}{j}$$ denote the $`k^{th}`$ harmonic number. Let $`\{S_n\}`$ be a random walk whose increments are equal $`k`$ with probability $`2/((k+1)(k+2))`$ for each integer $`k1`$. Let $`S:=S_{G1}`$ where $`G`$ is an independent geometric random variable with mean 2. Let $`\mathrm{\Theta }`$ be a random variable satisfying $$(\mathrm{\Theta }j)=1F_\mathrm{\Theta }(j1)=\underset{k=1}{\overset{\mathrm{}}{}}\frac{1}{k}\frac{1}{k+j}=\frac{H_j}{j}.$$ Assume $`\{S_n\}`$ and $`\mathrm{\Theta }`$ are independent of each other and of $`\{_t\}`$; denote expectation with respect to $`_𝐱`$ by $`𝔼_𝐱`$ and let $`𝔼`$ denote expectation with respect to the laws of $`S`$ and $`\mathrm{\Theta }`$. Analogously with $`\text{zeros}(𝐱)`$ we define the quantity $$\text{zeros}^{}(𝐱):=1+inf\{j>\text{ones}(𝐱):𝐱(\text{ones}(𝐱)1+j)=1\}$$ to be the number of zeros after the first block of ones; thus $`\text{zeros}^{}(𝐱)1`$, $`\text{zeros}^{}(𝐱)=\text{zeros}(𝐱)`$ if and only if $`\text{zeros}(x)1`$ and $`\text{zeros}(𝐱)=0`$ if and only if $`\text{ones}(𝐱)2`$. ###### Lemma 4 For any $`𝐱\text{ones}_j`$, $$𝔼_𝐱\text{jump}_1=\underset{k=1}{\overset{j}{}}\frac{1}{k}.$$ Equivalently, for any $`𝐱\mathrm{\Xi }`$, $$𝔼_𝐱\text{jump}_1=H_{\text{ones}(𝐱)}.$$ ###### Lemma 5 For any $`j1`$, any $`𝐱\text{zeros}_j`$, and any integer $`L1`$, $$_𝐱(\text{ones}(Y_1)L)(S+jL).$$ When $`\text{zeros}(𝐱)=0`$ then $$_𝐱(\text{ones}(Y_1)L)(\mathrm{\Theta }+\text{zeros}^{}(𝐱)B+SL)$$ where $`B`$ is a Bernoulli with mean $`1/2`$, and $`\mathrm{\Theta },B`$ and $`S`$ are all independent. Since $`S+\mathrm{\Theta }+\text{zeros}^{}(𝐱)`$ is an upper bound for both quantities $`S+\text{zeros}^{}(𝐱)`$ and $`S+\mathrm{\Theta }+B\text{zeros}^{}(𝐱)`$ appearing as stochastic upper bounds in Lemma 5, and since $`H_n`$ increases in $`n`$, we may put this together with Lemma 4 to obtain ###### Corollary 6 For any $`𝐱`$, $$𝔼_𝐱\text{jump}_2𝔼H_{\mathrm{\Theta }+\text{zeros}^{}(𝐱)+S}.$$ $`\mathrm{}`$ ###### Lemma 7 The conditional distribution of $`\text{zeros}^{}(Y_n)`$ given $`_{\sigma _{n1}}`$ is always bounded above stochastically by the law of $`\mathrm{\Theta }`$. In other words, for all $`j`$, $$(\text{zeros}^{}(Y_n)>j|_{\sigma _{n1}})1F_\mathrm{\Theta }(j).$$ Proof of Theorem 2 from the lemmas: It suffices to show that for any $`𝐱\mathrm{\Xi }`$, $$𝔼_𝐱\text{jump}_32.92.$$ We simply iterate conditional expectations and compute. By the Markov property, and Corollary 6, $`𝔼_𝐱\text{jump}_3`$ $`=`$ $`𝔼_𝐱𝔼_{Y_1}\text{jump}_2`$ $``$ $`𝔼_𝐱\left(𝔼_yH_{𝒮+\mathrm{\Theta }+\text{zeros}^{}(y)}\right)|_{y=Y_1}.`$ Since $`H_n`$ is increasing in $`n`$ we may use the stochastic upper bound in Lemma 7 for any $`𝐱`$ to see that this is at most $`𝔼H_{S+\mathrm{\Theta }^{(2)}}`$ where $`\mathrm{\Theta }^{(2)}`$ is the sum of two independent copies of $`\mathrm{\Theta }`$. The upper bound in Theorem 2 is completed by computing an upper bound for this. The function $`H`$ is concave and $`\mathrm{\Theta }^{(2)}1`$, so $$H(\mathrm{\Theta }^{(2)}+j)H(\mathrm{\Theta }^{(2)})H(j+1)1$$ and we may therefore conclude that $$𝔼H_{\mathrm{\Theta }^{(2)}+S}𝔼H_{\mathrm{\Theta }^{(2)}}+𝔼H_{S+1}1.$$ (3) To compute the quantity $`𝔼H_{\mathrm{\Theta }^{(2)}}`$, let $`\mathrm{\Theta }_1`$ and $`\mathrm{\Theta }_2`$ independently have the distribution of $`\mathrm{\Theta }`$ and write $`𝔼H_{\mathrm{\Theta }_1+\mathrm{\Theta }_2}`$ $`=`$ $`1+{\displaystyle \underset{1j,k}{}}{\displaystyle \frac{1}{j+k}}(\mathrm{\Theta }_1j)(\mathrm{\Theta }_2=k)`$ $`=`$ $`1+{\displaystyle \underset{1j,k}{}}{\displaystyle \frac{1}{j+k}}{\displaystyle \frac{H_j}{j}}\left({\displaystyle \frac{H_k}{k}}{\displaystyle \frac{H_{k+1}}{k+1}}\right)`$ $`=`$ $`1+{\displaystyle \underset{1j,k}{}}{\displaystyle \frac{1}{j+k}}{\displaystyle \frac{H_j}{j}}\left({\displaystyle \frac{H_k}{k}}{\displaystyle \frac{H_{k+1}}{k}}+{\displaystyle \frac{H_{k+1}}{k}}{\displaystyle \frac{H_{k+1}}{k+1}}\right)`$ $`=`$ $`1+{\displaystyle \underset{1j,k}{}}{\displaystyle \frac{1}{j+k}}{\displaystyle \frac{H_j}{j}}{\displaystyle \frac{H_{k+1}1}{k(k+1)}}.`$ We may evaluate this numerically as just under 2. For the last term on the RHS of (3), let $`\varphi (z)=𝔼z^S=_{n=0}^{\mathrm{}}z^n(S=n)`$ be the generating function for $`S`$, so $`z\varphi `$ is the generating function for $`S+1`$. For any positive integer $`j`$ there is an identity $$_0^1\frac{1z^j}{1z}𝑑z=_0^1(1+\mathrm{}+z^{j1})𝑑z=H_j.$$ Consequently, we may write $$𝔼H_{S+1}1=_0^1\frac{z𝔼z^{S+1}}{1z}𝑑z=_0^1z\frac{1\varphi (z)}{1z}𝑑z.$$ (4) To compute the generating function $`\varphi `$, first compute the generating function $`f`$ for the increments of $`\{S_n\}`$: $$f(z)=\underset{k=1}{\overset{\mathrm{}}{}}\frac{2}{(k+1)(k+2)}z^k=\frac{2zz^22(1z)\mathrm{log}\frac{1}{1z}}{z^2}.$$ Then $`\varphi =1/(2f)`$ and the integral in (4) becomes $$𝔼H_{S+1}=_0^1\frac{2z(\mathrm{log}\frac{1}{1z}z)}{3z^22z+2(1z)\mathrm{log}\frac{1}{1z}}𝑑z.$$ One may evaluate this numerically to approximately $`0.918797`$. Adding this to the value for $`𝔼H_{\mathrm{\Theta }^{(2)}}`$ gives a little under $`2.92`$. Rigorous bounds may be obtained with more care. $`\mathrm{}`$ ## 3 Proofs of Lemmas Proof of Lemma 4: By definition, each $`𝐱\text{ones}_j`$ begins with $`j`$ 1’s in positions $`0,\mathrm{},j1`$ followed by a zero. The evolution of $`\{M_t\}`$ decreases the binary representation $`_k2^k𝐱(k)`$, whence $`M_{\text{jump}_1^{}}\text{ones}_k`$ for some $`kj`$, that is, there is always a zero in some position in $`[0,j]`$. Furthermore, once there is a zero in position $`k`$ for some $`k<j`$, then there is always a zero at or to the left of position $`k`$. It follows that $`\text{jump}_1`$ is equal to the least $`k<j`$ for which $`\xi _{j,1}<\xi _{0,1}`$, that is, for which the clock in position $`k`$ goes off before the clock at position 0. The minimum is taken to be $`j`$ if there is no such $`k`$. It follows that $`_𝐱(\text{jump}_1=j)=1/j`$ and that for $`0<k<j`$, $$_𝐱(\text{jump}_1=j)=\frac{1}{k(k+1)}.$$ (5) To see this, note that $`\text{jump}_1=k`$ if and only if $`\xi _{0,1}`$ is the minimum of the exchangeable variables $`\{\xi _{k,1}:0k<j\}`$. Similarly, for $`0<k<j`$, $`\text{jump}_1=k`$ if and only if $`\xi _{k,1}`$ is the minimum of the variables $`\{\xi _{r,1}:0rk\}`$ and $`\xi _{0,1}`$ is the next least of the values. Computing expectations via (5) proves the lemma. $`\mathrm{}`$ Proof of Lemma 7: Again let us denote $`q=\text{ones}(𝐱)`$. We recall that $`(\text{jump}_1=k)=1/q`$ for $`k=q`$ and $`1/(k(k+1))`$ for $`1kq1`$. We claim that for any $`kq`$, $$(\text{zeros}^{}(Y_1)j|\text{jump}_1=k)\frac{k+1}{k+j}.$$ (6) If we can show this, then we will have $$(\text{zeros}^{}(Y_1)j)\frac{1}{q}\frac{q+1}{q+j}+\underset{k=1}{\overset{q1}{}}\frac{1}{k(k+1)}\frac{k+1}{k+j}.$$ Changing $`q`$ to $`q+1`$ increases this by $$\frac{j1}{(q+1)(q+j)(q+j+1)}$$ which is nonnegative. Setting $`q=\mathrm{}`$ then yields the upper bound in the lemma, and it remains to show (6). Observe first that it suffices to show this for $`k=q`$. This is because when $`k<q`$, the event $`\{\text{jump}_1=k\}`$ necessitates $`\xi _{k,1}=\mathrm{min}\{\xi _{r,1}:0rk\}`$. Thus to evaluate $`(\text{zeros}^{}(Y_1)j|\text{jump}_1=k)`$ we may wait until time $`\xi _{k,1}`$, at which point if no bit to the left of $`k`$ has flipped yet, the new conditional probability $`(\text{zeros}^{}(Y_1)j|_{\xi _{k,1}},\text{jump}_1=k)`$ is always at most $`(k+1)/(k+j)`$ by applying the claim for $`q=k`$. Assuming now that $`k=q`$, we note that the event $`\{\text{jump}_1=k\}`$ that we are conditioning on is just the event $$A:=\{\xi _{0,1}=\underset{0ik1}{\mathrm{min}}\xi _{i,1}\}$$ is the event that the clock at 0 goes off before any clock in positions $`1,\mathrm{},k1`$. Conditioning on $`A`$ then makes the law of $`\xi _{0,1}`$ an exponential of mean $`1/k`$ without affecting the joint distribution of $`\{\xi _{r,s}:r>k\}`$. Now let $`m_1`$ be the position at time $`\xi _{k,1}`$ of the first 1 to the right of $`k`$, and let $`t_1`$ be the time this 1 flips. Inductively, define $`m_{r+1}`$ to be the position of the first 1 to the right of $`k`$ after time $`t_r`$ and let $`t_{r+1}`$ be the first time after $`t_r`$ that this 1 flips. If the positions $`m_r,\mathrm{},m_r+j1`$ are not filled with ones at time $`t_{r1}`$ (define $`t_0=0`$) then it is not possible to have $`\text{zeros}^{}(Y_1)j`$ and $`A`$ and $`t_{r1}<\xi _{0,1}<t_r`$. That is, one cannot get from fewer than $`j`$ ones in the first block of ones to the right of $`k`$ to at least $`j`$ ones at the time of the flip at 0 without having the leftmost one in this block flip. On the other hand, if these $`j`$ positions are filled with ones at time $`t_{r1}`$, then $$(\text{zeros}^{}(Y_1)j,\xi _{0,1}<t_r|A,_{t_{r1}},\xi _{0,1}>t_{r1})\frac{k}{k+j}$$ since the event $`\{\text{zeros}^{}(Y_1)j,\xi _{0,1}<t_r\}`$ requires that the clock at 0 go off before the clocks in positions $`m_r,\mathrm{},m_r+j1`$ (recall that conditioning on $`A`$ has elevated the rate of the clock at 0 to rate $`k`$). Similarly, $`(\xi _{0,1}<t_r|A,_{t_{r1}},\xi _{0,1}>t_{r1})=k/(k+1)`$. Therefore, $$\frac{(\text{zeros}^{}(Y_1)j,\xi _{0,1}<t_r|A,_{t_{r1}},\xi _{0,1}>t_{r1})}{(\xi _{0,1}<t_r|A,_{t_{r1}},\xi _{0,1}>t_{r1})}\frac{k+1}{k+j}.$$ By the craps principle, the RHS is then an upper bound for the probability of $`\text{zeros}^{}(Y_1)j`$ conditioned only on $`\text{jump}_1=k`$. $`\mathrm{}`$ Proof of Lemma 5: Let $`𝐱\text{zeros}_j`$ and first assume $`j1`$. We prove the statement for $`n=1`$, the proof for greater $`n`$ being identical, conditioned on $`_{\sigma _{n1}}`$. It is simple to check whether $`\text{ones}(Y_1)=j`$. The bits in positions $`1,\mathrm{},j`$ will remain 0’s until the leading 1 flips at time $`\sigma _1`$, so the only thing to check is whether $`\sigma _1=\xi _{0,1}`$ is less than or greater than $`\xi _{j+1,1}`$. With probability $`1/2`$, $`\xi _{0,1}<\xi _{j+1,1}`$ and in exactly this case $`\text{ones}(Y_1)=j`$. Now condition on this inequality going the other way: $`\xi _{01}>\xi _{j+1,1}`$. Let $`t_1:=\xi _{j,1}`$. Let $`j+1+k_1`$ be the position of the first 0 of $`𝐱`$ to the right of position $`j+1`$. Then at time $`t_1`$, the position of the first one to the right of $`j+1`$ is $`j+Z_1`$, where $`Z_1`$ is the least $`l[1,k_11]`$ for which $`\xi _{j+1+l,1}<\xi _{j+1,1}`$. If no such $`l`$ exists, then $`Z_1=k_1`$. We compute $`_𝐱(Z_1=l)`$ as follows. The variables $`\{\xi _{r,1}:r\{0\}[j+1,j+k_1]\}`$ are independent exponentials. For $`1l<k_1`$, the event that $`\text{ones}(Y_1)j`$ and $`Z_1=l`$ is the intersection of the event $`A`$ that $`\xi _{j+1,1}`$ is less than $`\xi _{0,1}`$ and $`\xi _{r,1}`$ for all $`j+2rj+l`$ with the event $`B`$ that $`\xi _{j+1+l,1}<\xi _{j+1,1}`$. In other words, among $`l+2`$ independent exponentials, the index of the least and second least must be $`j+2`$ and $`j+1`$ respectively. The unconditional probability of this is $`1/((l+1)(l+2))`$. Having conditioned on the larger event $`\{\xi _{j+1,1}<\xi _{0,1}\}`$, the conditional probability is therefore equal to $`2/((l+1)(l+2))`$. This holds for $`l<k_1`$, where $`Z_1=k_1`$ with the complementary probability. To sum up, $`Z_1`$ is distributed as $`S_1k_1`$ where $`S_1`$ has the distribution of the random walk increments described in the lemma. The last step is to invoke the Markov property. Condition on $`_{t_1}`$. The chain from here evolves under the law $`_{X(t_1)}`$. Iterating the previous argument, there are two cases. The first case, which happens with probability $`1/2`$ is that the clock at the origin goes off before the next alarm at location $`j+1+Z_1`$. In this case, $`\text{ones}(Y_1)=j+Z_1`$. In the alternative case, we let $`t_2`$ be the time at which the clock at location $`j+1+Z_1`$ next goes off. We let $`Z_2`$ be the number of consecutive 1’s at time $`t_2^{}`$ starting from position $`j+1+Z_1`$. Then conditional on $`_{t_1}`$, $`Z_2`$ is distributed as $`S_1k_2`$ where $`k_2`$ is the number of consecutive 1’s at time $`t_1`$ starting at position $`j+1+Z_1`$. Iterating in this way, we have the following inductive definitions. Let $`t_0=0`$. Let $`\tau `$ be the least $`r`$ for which the clock at the origin goes off after time $`t_r`$ but before the first alarm at location $`j+1+_{i=1}^rZ_i`$. For each $`r\tau `$, we may define $`k_r`$ to be the number of consecutive 1’s at time $`t_{r1}`$ starting at location $`j+1+_{i=1}^{r1}Z_i`$. We may then define $`t_r`$ to be the first time after $`t_{r1}`$ that the alarm at location $`j+1+_{i=1}^{r1}Z_i`$ goes off, and we may define $`Z_r`$ so that $`j+1+_{i=1}^rZ_i`$ is the location of the first zero to the right of $`j+1+_{i=1}^{r1}Z_i`$ at time $`t_r^{}`$. The upshot of all of this is that $$\text{ones}(Y_1)=j+\underset{i=1}{\overset{\tau }{}}Z_i$$ and that the joint distributions of $`\tau `$ and $`\{Z_i:1i\tau \}`$ are easily described. Conditioned on $`\tau r`$ and on $`_{t_r}`$, the probability of $`\tau =r+1`$ is always $`1/2`$; as well, $`Z_{r+1}`$ given $`\tau r+1`$ and $`_{t_r}`$ is always distributed as a truncation of $`S_1`$. We conclude that $`\text{ones}(Y_1)`$ is stochastically dominated by the sum of $`\tau `$ independent copies of $`S_1`$, hence as $`S_{G1}`$. Finally, we consider the case where $`\text{zeros}^{}(𝐱)=l>\text{zeros}(𝐱)=0`$. Let $`q=\text{ones}(𝐱)`$, so that $`𝐱`$ begins with $`q`$ ones, followed by $`l`$ zeros, followed by a one in position $`q+l`$. A preliminary observation is that if we begin with a one at the origin, the position $`W(t)`$ of the leading one at a later time $`t`$ is an increasing function of $`t`$; hence, if $`T_\mu `$ is an independent exponential with mean $`\mu `$, the distribution of $`W(T_\mu )`$ is stochastically increasing in $`\mu `$. Begin by writing $$_𝐱(\text{ones}(Y_1)j)=\underset{k=1}{\overset{q}{}}(\text{ones}(Y_1)j,\text{jump}_1=k).$$ Let $`l^{}`$ denote the number of zeros consecutively starting from position $`k`$ at time $`\xi _{k,1}`$ if $`\text{jump}_1=k<q`$, and $`l^{}=l`$ if $`k=q`$. In other words, $`l^{}=\text{zeros}^{}(𝐱^{})`$ where $`𝐱^{}`$ is the word at the last time $`t`$ that ones changes before the leading bit flips ($`t=\xi _{\text{jump}_1,1}`$ if $`\text{jump}_1<q`$ and $`t=0`$ otherwise). We may then describe $`\text{ones}(Y_1)`$ as $`l^{}+W`$, where $`W`$ is the number of consecutive positions starting from position $`\text{jump}_1+l^{}`$ that turn to zeros between time $`t`$ and $`\xi _{0,1}`$. Now we break into two cases. Condition first on $`\{\text{jump}_1=q\}`$. The time $`\xi _{0,1}`$ is now an exponential of mean $`1/q`$, and before this time, the bits from position $`q+l`$ onward evolve independently. We may describe $`\text{ones}(Y_1)`$ as $`l+W(\xi _{0,1})`$, where $`W`$ is the number of consecutive positions starting at $`q+l`$ which have become zeros in the time from 0 to $`\xi _{0,1}`$. The first part of this lemma established that when $`\xi `$ has rate 1, then $`W(\xi )S`$. Our preliminary observation now shows, conditional on $`\{\text{jump}_1=q\}`$, that $`W(\xi _{0,1})S)`$. Next, let us condition on $`\text{jump}_1=k<q`$, obeserving that then $`l^{}qk`$. In order to have $`l^{}r`$, it is necessary that $`\xi _{k,1}=\mathrm{min}\{\xi _{s,1}:ksk+r1\}`$. Having conditioned on $`\text{jump}_1=k`$, the distribution of $`\xi _{k,1}`$ becomes an exponential of rate $`k+1`$, so that the conditional probability of this clock going off before $`r1`$ other conditionally independent clocks of rate 1 is just $`(k+1)/(k+r)`$. Since the event $`\{\text{jump}_1=k<q\}`$ has probability $`1/(k(k+1))`$ if $`k<q`$ and zero otherwise, we may remove the conditioning and sum to get $$(l^{}r,\text{jump}_1<q)\underset{k=1}{\overset{q1}{}}\frac{1}{k(k+1)}\frac{k+1}{k+r}(\theta ^{}r).$$ We also still have in this case $`WS`$. Putting together the cases $`\text{jump}_1=q`$ and $`\text{jump}_1<q`$, we see that $`l^{}=l`$ with probability $`1/q`$ and otherwise $`l^{}\mathrm{\Theta }`$. The crude bound $`1/q1/2`$ gives $`l^{}\mathrm{\Theta }+\text{zeros}^{}(𝐱)B`$. Since $`l^{}\sigma (_t)`$ and the bound $`WS`$ holds conditionally on $`_t`$, we arrive at $`\text{ones}(Y_1)\mathrm{\Theta }+\text{zeros}^{}(𝐱)B+S`$. $`\mathrm{}`$ ## 4 Lower bound To arrive at a lower bound, consider a tree whose vertices are positive integers, identified with their binary expansions. The root is 1, and the children of $`x`$ are $`2x`$ and $`2x+1`$. Associated with each node $`x`$ are a set of $`n(x)`$ possible transitions, where $`n(x)`$ is the number of 1’s in the binary expansion of $`x`$. The transitions are to numbers gotten by flipping a 1 and simultaneously flipping all bits to its right. Note that all transitions from $`x`$ are to numbers less than $`x`$. Let $`r(x)`$ denote the reward if the leading bit of $`x`$ is flipped, namely one less than the number of leading 1’s in the binary expansion of $`x`$. Recursively, we assign to each $`x`$ a mean reward and maturation time, $`(a(x),b(x))`$ as follows. Fix a set $`B`$ of bad nodes, to be determined later. On first reading, take $`B`$ to be empty. For any $`x`$, let $`y_1,\mathrm{},y_n`$ be the possible transitions from $`x`$. Let $`(a(x),b(x)):=(0,0)`$ if $`xB`$ and otherwise, let $$(a(x),b(x)):=\frac{1}{n}\left((r(x),1)+\underset{i=1}{\overset{n}{}}(a(y_i),b(y_i))\right).$$ Let $`𝒯`$ be any finite binary rooted subtree, meaning that any vertex in the subtree has either zero or two children in the subtree. ###### Lemma 8 Suppose no leaf of $`𝒯`$ is in $`B`$. Then an almost sure lower bound for the lim inf speed from any starting configuration is given by the minimum over leaves $`x`$ of $`T`$ of $`1+a(x)/b(x)`$. The lower bound in Theorem 2 will follow from Lemma 8 together with an implementation of the recursion. Below is some code written in C that implements the recursion for a complete binary tree of depth 15, with the set B chosen to give a good bound without much trouble. ``` #include <stdio.h> #define N 18 #define M 524288 main() { float a[M][2]; int i , j , k, n, min_index, locmin, power[N+2], bitcount=0, lastcount=0; Ψfloat min_ratio; float locrat; float average; a[0][0] = 0; a[0][1]=0; a[1][0] = 0; a[1][1] = 0; power[0] = 1; power[1]=2; average = 0; /* the n loop stratifies the recursion by levels of the tree */ for (n=1; n <= N; n++) { power[n+1] = 2*power[n]; /* the i loop indexes vertices inside a level */ for (i=power[n]; i < power[n+1]; i++) { bitcount = 0; lastcount=0; a[i][0] = 0; a[i][1] = 0; for (j=0; j <= n; j++) { k = i^(power[j]); if (k < i) { /* k < i iff there is a 1 in position k */ k = power[j+1]*(i / (power[j+1])) + power[j+1] - 1 - i % power[j+1]; a[i][0] += a[k][0]; a[i][1] += a[k][1]; bitcount++; Ψ} else lastcount=bitcount; /* lastcount will eventually say how many 1’s after the leftmost zero */ } a[i][1]++; a[i][0] += bitcount - lastcount - 1; a[i][0] = a[i][0] / bitcount; a[i][1] = a[i][1] / bitcount; /* a particular choice of the set B is given in literal form */ if ( i==2 || i==4 || i==5 || i==8 || i==9 || i==10 || i==11 || i==18 || i==19 || i==20 || i==21 || i==22 || i==23 || i==36 || i==37 || i==38 || i==40 || i==41 || i==42 || i==43 || i==44 || i==45 || i==46 || i==47 || i==73 || i==74|| i==75 || i==80 || i==81 || i==82 || i==83 || i==84 || i==85 || i==86 || i==87 || i==88 || i==89 || i==90 || i==91 || i==160 || i==161 || i==162 || i==163 || i==164 || i==165 || i==166 || i==167 || Ψ i==168 || i==169 || i==170 || i==171 || i==172 || i==173 || i==174 || i==175 || i==178 || Ψ i==180 || i==181 || i==182 || i==183 || i==324 || i==325 || i==326 || i==328 || i==329 || i==330 || i==331 || i==332 || i==333 || i==334 || i==335 || i==336 || i==337 || i==338 || i==339 || i==340 || i==341 || i==342 || i==343 || i==344 || i==345 || i==346 || i==347 || i==348 || i==349 || i==350 || i==351 || i==361 || i==362 || i==363 || i==656 || i==657 || i==658 || i==659 || i==660 || i==661 || i==662 || i==663 || Ψ i==672 || i==673 || i==674 || i==675 || i==676 || i==677 || i==678 || i==679 || i==680 || i==681 || i==682 || i==683 || i==684 || i==685 || i==686 || i==687 || i==688 || i==689 || i==690 || i==691 || i==692 || i==693 || i==694 || i==695 || i==697 || i==698 || i==699|| i==1322 || i==1323 || i==1346 || i==1347 || i==1348 || i==1349 || i==1350 || i==1351 || i==1352 || i==1353 || i==1354|| i==1355 || i==1356 || i==1357 || i==1358 || i==1359 || i==1360 || i==1361 || i==1362 || i==1363 || i==1364 || i==1365 || i==1366 || i==1367 || i==1368 || i==1369 || i==1370 || i==1371 || i==1372 || i==1373 || i==1374 || i==1375 || i==1380 || i==1381 || i==1384 || i==1385 || i==1386 || i==1387 || i==1388 || i==1389 || i==1390 || i==1391 || i==2704 || i==2705 || i==2706 || i==2707 || i==2708 || i==2709 || i==2710 || i==2711 || i==2720 || i==2721 || i==2722 || i==2723 || i==2724 || i==2725 || i==2726 || i==2727 || i==2728 || i==2729 || i==2730 || i==2731 || i==2732 || i==2733 || i==2734 || i==2735 || i==2736 || i==2737 || i==2738 || i==2739 || i==2740 || i==2741 || i==2742 || i==2743 || i==2745 || i==2746 || i==2747 || i==5456 || i==5457 || i==5458 || i==5459 || i==5467 || i==5460 || i==5461 || i==5462 || i==5463 || i==5465 || i==5466 ) Ψ { ΨΨa[i][0] = 0; a[i][1] = 0; Ψ } if(i == (M/2)) { Ψmin_index = i; Ψmin_ratio = a[i][0]/a[i][1]; } if(i > (M/2)) { Ψif((a[i][0]/a[i][1]) < min_ratio) Ψ { Ψ min_index = i; Ψ min_ratio = a[i][0]/a[i][1]; Ψ } } if (i > (M/2)) Ψ { Ψ average = average + (a[i][0]/a[i][1]); } if(i == 3) { Ψlocmin = i; Ψlocrat = a[i][0]/a[i][1]; } if(i < (M/2)) Ψ{ if (i > 71) { Ψ if(a[i][1] > 0) { if ((a[i][0]/a[i][1]) < locrat) Ψ { Ψ locmin = i; Ψ locrat = (a[i][0])/(a[i][1]); Ψ } Ψ } } } } printf("\n"); } printf("Row with minimum ratio is:\n"); printf("min_index=%i,exp. reward=%f, exp. time=%f, ratio=%f\n",min_index, a[min_index][0], a[min_index][1], a[min_index][0]/a[min_index][1]); printf("Row with local minimum ratio is:\n"); printf("locmin=%i,exp. reward=%f, exp. time=%f, ratio=%f\n", locmin, locrat, average, M); Ψ}} ``` A look at the data shows the minimum value of $`a(x)/b(x)`$ on each level to be obtained when the binary expansion of $`x`$ alternates. In particular, the global minimum is at $`x=349525`$ and has value $`0.646\mathrm{}`$, which proves the lower bound. Proof of Lemma 8: Any finite rooted binary subtree induces a prefix rule, that is, a map $`\eta `$ from infinite sequences beginning with a 1 to leaves of $`𝒯`$, defined by $`\eta (𝐱)=w`$ for the unique leaf of $`𝒯`$ that is a prefix of $`\eta `$. Given a trajectory of the Markov chain $`\{X_t\}`$, define a sequence of elements of $`𝒯`$ as follows. Let $`x_0:=\eta (X_0)`$ be the prefix of the initial state of the trajectory. Let $`\tau _0:=0`$. As the definition proceeds, verify inductively that for $`\tau _kt<\tau _{k+1}`$, the string $`x_k`$ will be an initial segment of $`X_t`$. The recursion is as follows. Let $`\tau _{k+1}`$ be the first time after $`\tau _k`$ that a 1 flips in the initial segment $`x_k`$ of $`X_t`$. Let $`x_{k+1}^{}`$ be the string gotten from $`x_k`$ by flipping this bit and all bits to its right. If $`x_{k+1}^{}B`$ and $`x_{k+1}^{}`$ is not the zero string then let $`x_{k+1}`$ be $`x_{k+1}^{}`$, stripped of any leading zeros. If $`x_{k+1}^{}B`$ or $`x_{k+1}^{}`$ is the zero string, then let $`x_{k+1}=\eta (X_{\tau _{k+1}})`$. Let $`\rho _0,\rho _1,\mathrm{}`$ be the successive times that this latter transition occurs, that is, successive times $`t`$ in the recursion that $`\eta (X_t)`$ is computed. Given $`x𝒯`$, suppose we begin counting every time the sequence $`\{x_k\}`$ hits $`x`$ and stop counting at every time $`\rho _k`$. When we are counting, we count how many times the leading 1 flips, and how many 1’s flip together when this happens, or more precisely, we keep a cumulative count, each time adding one less than the number of 1’s that have flipped together. Let this cumulative count be denoted $`(A(x,t),B(x,t))`$, where $`A`$ counts flips of the leading 1. If $`xB`$, then by convention we do not count anything. Claim: For each $`x𝒯B`$, if $`x`$ is visited infinitely often then the ratio of $`A(x,t)/B(x,t)`$ converges as $`t\mathrm{}`$ to $`a(x)/b(x)`$. Proof: Conditioned on the past, each time we start counting we are equally likely to transition to each of the $`n(x)`$ possible transitions. The claim then follows from induction on $`x`$. Each $`x`$ defines a set of time where $`x`$ is on, that is, where we are counting leading 1 flips and rewards for that $`x`$. As $`x`$ ranges over the leaves of $`𝒯`$, the on-timesets for $`x`$ partition $`[0,\mathrm{})`$. In each such time set, we have shown that the mean number of recorded simultaneous 1-flips per leading 1 flip is $`1+a(x)/b(x)`$ in the limit, as long as the time set is unbounded. Since the partition is finite, it follows that the lim inf speed is at least the minimum of the values. $`\mathrm{}`$ We remark that the only place this computation is not sharp is when the number of 1’s flipping together exceeds the number recorded, because the present knowledge of the prefix was a string of all 1’s and there were more 1’s after this that also flipped. Thus by making the tree $`𝒯`$ big enough, even without increasing $`B`$, we can get arbitrarily close to the true value. ## 5 Further observations The following argument almost solves Problem (3a), and perhaps may be strengthened to a proof. Lemma 4 of \[GHZ98\] is proved by means of a duality result. The result is that the probability, starting from a uniform random state, of finding a 1 in position $`r`$ after $`t`$ steps (counting suppressed transitions), is equal to half the probability that $`𝐱^r`$ has not reached the minimum yet after $`t`$ steps (again counting suppressed transitions). The argument that proves this may be generalized by introducing a simultaneous coupling of the process from all starting states. The probability, from a uniform starting state, of finding a 1 in every position in a set $`A`$ after $`t`$ steps, is then the expectation of the function that is zero if the column vector of all 1’s is not in the span of the columns of the matrix whose rows are the states reached at time $`t`$ starting at $`𝐱^r`$, as $`r`$ varies over $`A`$, and is $`2^u`$ if the column vector of all 1’s is in the span and the matrix has rank $`u`$. The kernel of the matrix is the set of starting configurations that reach the minimum by time $`t`$ (the simultaneous coupling is linear). Hence, as long as $`A`$ and $`t`$ are such that the probability of reaching the minimum from any $`𝐱^r`$ by time $`t`$ goes to zero, the rank of the matrix will be $`|A|`$ and the probability of finding all 1’s in positions in $`A`$ at time $`t`$ will go to $`2^{|A|}`$. In particular, if a window of fixed size moves rightward faster than the limsup speed, then what one sees in this window approaches uniformity. This is not good enough to imply uniformity of a window a fixed distance to the right of the leftmost 1. Acknowledgements: We would like to say thanks to Attila Pór, sharing this problem, and to Jiri Matousek giving references about the Klee-Minty cube.
warning/0506/astro-ph0506769.html
ar5iv
text
# The large asymmetric HI envelope of the isolated galaxy NGC 864 (CIG 96) ## 1 Introduction The origin of asymmetries in isolated galaxies is not well understood. Are they always induced by the presence of small companions? What is the influence of recent captures, and how long does it take for a parent galaxy to become again axisymmetric afterwards? Could some originate from internal, as yet not well studied, long-lived dynamical instabilities (e.g. Baldwin et al. 1980)? Integral HI spectra provide a powerful statistical approach to this issue since they contain information about both the HI density distribution and the velocity field. We have observed, or compiled from the literature, single-dish 21-cm line spectra for almost 800 galaxies (Espada et al. in prep.) as part of the AMIGA project (Analysis of the interstellar Medium of Isolated GAlaxies; Verdes-Montenegro et al. 2005), which aims to provide a baseline against which environmental effects can be evaluated. We have constructed a complete sample of the most isolated galaxies, starting from the Catalogue of Isolated Galaxies (CIG, Karachentseva 1973; 1050 galaxies), selected on the basis of the distance to the nearest similarly sized galaxies. About 50% of these galaxies show asymmetric HI profiles, consistent with the results from previous work on smaller samples of isolated galaxies (104 galaxies in Haynes et al. 1998; 30 galaxies in Matthews et al. 1998). Surprisingly, other samples of galaxies in denser environments show only slightly larger values of this rate (50 - 80%, e.g. Swaters et al. 2002, Richter & Sancisi 1994, Sulentic & Arp 1983). HI synthesis imaging of very isolated galaxies with a highly lopsided HI profile may clarify the origin of these asymmetries. Not only HI-rich companions can be identified, but the presence of tidal features can be revealed, which would trace a past interaction. We have selected from our database a well defined sample of the most isolated galaxies showing significant asymmetries in their HI profiles (A<sub>n</sub> $`>`$1.1, see Haynes et al. 1998 for the definition), and mapped them with the VLA<sup>1</sup><sup>1</sup>1The National Radio Astronomy Observatory is a facility of the National Science Fundation Operated under cooperative agreement by Associated Universities, Inc. in the 21-cm HI line in order to look for signs of external interaction by analyzing in detail the HI distribution and its kinematics. One of the most interesting galaxies in our VLA sample is NGC 864, number 96 in the CIG catalog, which, as we shall see, shows a peculiar HI envelope in our 21-cm maps. It is classified as SAB(rs)c by de Vaucouleurs et al. (1991). It has an optical size of 4.7 x 3$`\text{.}^{}`$5 and an apparent magnitude m<sub>B</sub> = 11.6. It has a very asymmetric HI profile (cf. Fig. 1). The systemic heliocentric velocity that we have measured for this galaxy (see Sec. 3) is 1561.6 km s<sup>-1</sup>, corresponding to a distance of 17.3 Mpc (H<sub>0</sub> = 75 km s<sup>-1</sup> Mpc<sup>-1</sup>) after applying the correction w.r.t. the CMB reference frame. This gives an optical luminosity of 1.1 $`\times `$ 10<sup>10</sup> L. The closest galaxy to NGC 864 at the same redshift is a small companion that we detected in our HI observations (see Sec. 2) with v<sub>hel</sub> = 1605 km s<sup>-1</sup>, located at a distance d $``$ 80 kpc (15$`\text{.}^{}`$3) of NGC 864, visible in the POSS2 as a faint object (identified in NED as 2MASX J02162657+0556038 with a reported magnitude of m<sub>B</sub> = 16.38) with an approximate size of 0$`\text{.}^{}`$8 $`\times `$ 0$`\text{.}^{}`$6 (physical diameter $``$ 4 kpc). The other galaxies within a projected radius of d = 0.5 Mpc from NGC 864 are fainter by 3 to 7 magnitudes. ## 2 HI synthesis observations and results Observations of the 21 cm line for NGC 864 were made with the VLA in its D configuration in July 2004 with 26 antennas. A bandwidth per IF of 3.125 MHz (from 1249.5 to 1895.2 km s<sup>-1</sup>) was used, in 2IF correlator mode, giving a velocity resolution of 48.8 kHz (10.4 km s<sup>-1</sup>) for the 64 individual channels after Hanning smoothing. The data were calibrated using the standard VLA procedure in AIPS and were imaged using IMAGR. The average of the line-free channels has been subtracted from all the individual channels. The galaxy was detected between 1436.7 km s<sup>-1</sup> and 1686.6 km s<sup>-1</sup>. The synthesized beam was 49$`\text{.}^{\prime \prime }`$8 $`\times `$ 46$`\text{.}^{\prime \prime }`$2 ($`\alpha \times \delta `$). The rms noise level achieved after 4 hours is 0.66 mJy/beam. Primary beam correction has been applied to our maps. In Fig. 4 .a) and Fig. 4 .b) we display the channel maps smoothed to a beam size of 70$`\text{.}^{\prime \prime }`$4 $`\times `$ 65$`\text{.}^{\prime \prime }`$3, containing the HI emission at the observed radial velocities, from 1426.3 to 1707.4 km s<sup>-1</sup>. We detect the small companion in the HI map from 1572.0 to 1655.3 km s<sup>-1</sup> at $`\alpha `$(2000.0) = 02<sup>h</sup>16<sup>m</sup>26$`\text{.}^{\prime \prime }`$9 and $`\delta `$(2000.0) = 055624$`\text{.}^{\prime \prime }`$0 . The total spectrum has been obtained by integrating the emission in the individual channel maps (solid line in Fig. 1, top). We have measured a HI content of M<sub>HI</sub> = 7.53 $`\times `$ 10<sup>9</sup> M, in very good agreement with the single dish profile obtained by Haynes et al. (1998, dotted line in Fig. 1, top), showing that there is no loss of flux in the synthesis imaging. The profile of the small companion is shown at the bottom of the same figure. Subtraction of the flux of this dwarf galaxy from the total spectrum will not change its (asymmetric) shape. In Fig. 2 we show the integrated HI column density distribution of NGC 864 (left) and the HI velocity field (right), both overlaid on a POSS2 red band image, the deepest one we found. These maps have been calculated as follows. The channel maps were smoothed with a gaussian tapering function, and then non-signal pixels of the maps were blanked out. The channel maps where emission was detected were added up to produce the integrated emission map. The intensity weighted mean radial velocity field has been obtained in a similar way. The small companion mentioned above is clearly seen, together with its faint optical counterpart. The integrated emission map shows an unresolved HI depression in the center of the galaxy, while the highest column densities are located in a pseudoring structure with a size of $``$ 90<sup>′′</sup> $`\times `$ 55<sup>′′</sup>, better seen in the grey scale map shown in Fig. 3. This seems to trace the outer spiral structure of NGC 864, although the spatial resolution of the HI data is too small for a detailed comparison. An arm-like feature seems to join this structure at a second enhancement in the column density distribution. This enhancement has a ring-like shape with an approximate size of 8$`\text{.}^{}`$2 $`\times `$ 4$`\text{.}^{}`$6 (Fig. 2 and 3), nearly doubling the size of the optical disk. It is more prominent NW and SE of the optical disk and slightly closed to the NE, while undetected to the SW. The receding side of NGC 864 is narrower than the approaching one. The velocity field shows a symmetric pattern of differential rotation in the inner parts (r $``$ 300<sup>′′</sup>), while the kinematics of the outer parts is strikingly different. The outer isovelocity contours in the northern part show some twisting in the E direction, the last contour being nearly closed as a sign of a flattening rotation curve, and the southern isovelocity contours bend to the south with decreasing values characteristic of a declining rotation curve. The optical disk is reasonably symmetric in both spiral structure and angular extent, except for a slight enhancement of the southern arm, and a minor extension of the optical disk to the SE. ## 3 Modeling of the galaxy We have modeled the velocity-field of NGC 864 with a least-square algorithm based on a tilted ring model (Begeman 1987; the ROTCUR task in GIPSY). By this procedure we divided the galaxy into concentric rings, each of them with a width of 20<sup>′′</sup> along the major axis. The asymmetry of the HI distribution and velocity field in the outer parts lead us to model separately the approaching and the receding part of the galaxy. Points within a sector of $`\pm `$ 30 from the minor axis were excluded from the fits. The center position was fixed to the position of the optical center $`\alpha `$(2000.0) = 02<sup>h</sup>15<sup>m</sup>27$`\text{.}^{\prime \prime }`$6 and $`\delta `$( 2000.0) = 060009$`\text{.}^{\prime \prime }`$1 (Leon & Verdes-Montenegro 2003). The systemic velocity of 1561.6 km s<sup>-1</sup> has been obtained as the central velocity of the HI spectrum at the 20% level. Expansion velocities were set to zero. In a first iteration the position angle, the inclination and rotation velocity have been left free, giving already a very satisfactory modeling of the data cube. The derived rotation curve, however, suffers from beam smoothing: since the beam is elongated along the major axis the integration effect biases the radial velocities towards values lower than the true values corresponding to the rotation curve. A second iteration has been performed, this time fixing the radial velocities at slightly higher values in order to correct for beam smearing, and leaving again the position angle and the inclination free. The modeled cubes for the redshifted and blueshifted parts of NGC 864 were combined in a single cube from which a velocity field has been obtained ( Fig. 5). It shows a remarkable agreement with the observed one (Fig. 2 right), although not reproducing the asymmetries along the minor axis direction. Deriving the position angle of NGC 864 from our modeling for r $``$ 300<sup>′′</sup> we obtain a mean value of 23 $`\pm `$ 3, in very good agreement with the optical disk orientation. The inclination obtained as the mean between the receding and approaching sides is 43 $`\pm `$ 2, which is consistent to the value given by Tully (1988) of 45. The modelled inclinations for both sides are almost constant and their values are always close to the mean. The major axis twists both in the northern and southern part of NGC 864, although is more pronounced to the north. There it starts at r $``$ 240<sup>′′</sup> at nearly the edge of the optical disk, reaching a position angle of 39. In the southern part it starts at r $``$ 360<sup>′′</sup> going up to 31. In Fig. 6 we compare the position-velocity cut at 23 for the modeled (top) and observed (bottom) cube. Again the observations are very well reproduced. The plot also shows the projected rotation curve used. This figure contains the essential peculiarities of NGC 864: asymmetric HI distribution and kinematics, with a noticeable drop in rotation velocities. We note that the blueshifted part of the observed position-velocity diagram shows a faster than keplerian drop in the velocity of NGC 864, with a trend to reach r<sup>-0.5</sup> values at larger radii, that should of course be explained by non-circular motions and/or projection effects. Such a phenomenon is extremely rare in HI envelopes. ## 4 Discussion and conclusions Our results for NGC 864 have wider implications : a) The integral HI profile is symmetric in velocity, but asymmetric in intensity. Yet the 2D-kinematics of NGC 864 show large scale asymmetries: there is a kinematic warp, and in the southern part there is an abrupt decline of the rotation velocity, as is evident in Fig. 6. The atomic gas here looks like a kinematically detached clump , evident as a secondary peak in the position velocity cut at radii $``$ 400<sup>′′</sup> to the SW. This region has a physical extent of around 3$`\times `$4 in the major axis and minor axis directions, respectively. Clearly a simple integrated profile analysis will conclude that the central parts of NGC 864 are roughly symmetric, with a value of $`\mathrm{\Delta }`$V<sub>20</sub> correlated with the luminosity of the galaxy, since it falls on the Tully-Fisher relation. It is the outer parts which are surprisingly asymmetric, but since the radial velocities in the approaching side are closer to the systemic velocity than those in the central parts, the asymmetry only manifests itself in a higher amplitude of the blueshifted horn of the HI profile. b) The outer HI envelope is large, massive, very asymmetric and presents much structure. Its HI mass is about 4.5 $`\times `$ 10<sup>9</sup> M, and the HI mass associated with the steep drop of the rotation curve is about 2.3 $`\times `$ 10<sup>8</sup> M. No optical counterpart is found associated to the perturbed atomic gas. Warped HI envelopes are common around spirals, but few have so much structure and asymmetries as the one reported here. c) The galaxy is isolated with respect to similarly sized galaxies by a rather strict criterion. Yet there are 5 small companions within a projected distance of 500 kpc. The closest one, detected in our maps, has a dimensionless gravitational interaction strength (Dahari 1984) of 4 $`\times `$ 10<sup>-4</sup>, while for the others this parameter ranges from 1 $`\times `$ 10<sup>-5</sup> to 8 $`\times `$ 10<sup>-7</sup>. They are hence too small in mass to have caused the strong perturbations in the outer envelope of NGC 864. A similar conclusion follows from considerations about spiral forcing by companions (cf. Athanassoula 1984). We have considered several possibilities to explain the origin of these HI asymmetries. A self-induced perturbation seems highly unlikely: while the pseudoring is a very standard feature, the outer HI ring-like structure is too large to be a broken resonance ring, since the outer Lindblad resonance is usually located a bit over 2 times the end of the bar (Athanassoula et al 1982). An external perturbation by a companion, which could have interacted or even been accreted, seems likely to have caused the SW clump and the outer ring-like structure. However, we can exclude an encounter with a large pericenter distance since this would necessitate a massive companion, which is simply not there. A small companion would have to come very near NGC 864 or, even better, go through it, but a central, or near-central, passage will (cf. Athanassoula et al. 1997, Berentzen et al. 2003) yield results that do not match at all the morphology of NGC 864. The alternative left is that the companion crossed the equatorial plane of the target at an intermediate distance, e.g. just outside the optical disc and still within the extended HI disc. Such a passage could have induced the warp, and if the intruder was a loosely bound gas rich dwarf, its gas could have contributed to the SW clump, which has a similar gas mass, while its stars and dark matter could have dispersed. To study this scenario, which is of course only a cartoon – and to explain the form of the ring-like density enhancement, the mass of the excess gas and the velocity perturbations – full blown selfconsistent simulations, including both gas and stars, are necessary, but beyond the scope of the present paper. ###### Acknowledgements. DE, SL and LVM are partially supported by DGI Grant AYA 2002-03338 and Junta de Andalucía TIC-114 (Spain). LV-M acknowledges the hospitality of the Observatoire de Marseille. We used the NASA/IPAC Extragalactic Database (NED), operated by the Jet Propulsion Laboratory, California Institute of Technology, under contract with the National Aeronautics and Space Administration.
warning/0506/astro-ph0506703.html
ar5iv
text
# Dynamical models of Kuiper Belt dust in the inner and outer Solar System ## 1 Introduction Kuiper Belt Objects (KBOs) are icy bodies that lie in a disk beyond Neptune’s orbit. It is estimated that there are about 10<sup>5</sup> objects with diameters greater than 100 km in the 30–50 AU annulus and a total population roughly 3 orders of magnitude larger than that of the asteroid belt (Jewitt & Luu 2000). The outer limit of the belt is presently not well determined but may be near 50 AU (Chiang & Brown 1999; Allen, Bernstein & Malhotra 2001). Stern (1996) suggested that mutual collisions among KBOs can generate debris at a rate of (0.0095–3.2)$`\times `$10<sup>11</sup> g s<sup>-1</sup>. Using this estimate, Yamamoto & Mukai (1998) calculated a dust production rate of (0.0086–2.9)$`\times 10^7`$ g s<sup>-1</sup> in particles smaller than 10 $`\mu `$m. The impacts of interstellar dust on KBOs is also a significant source of interplanetary dust particles. Yamamoto & Mukai (1998) estimated that if there are $``$10<sup>13</sup> KBOs of radius $``$0.1 km, the total dust production rate for particles smaller than 10$`\mu `$m is (0.37–2.4)$`\times 10^6`$ g s<sup>-1</sup> if the objects have hard icy surfaces, or (0.85–3.1)$`\times 10^7`$ g s<sup>-1</sup> if the objects are covered with icy particles smaller than the interstellar grain impactors. Detectors on board Pioneer 10 and 11 have indeed been able to detect in situ dust in the outer solar system (see Landgraf et al. 2002). Interstellar dust grains are estimated to contribute less than one percent to the measured flux, therefore this dust is thought to have a solar system origin. The dynamical signatures indicate that the dust detected between Jupiter and Saturn is due to short period external Jupiter family comets and short period Oort cloud comets. The dust detected beyond 10 AU (outside Saturn’s orbit) is most likely produced by KBOs. If so, then a KB dust production rate of 2$`\times `$10<sup>14</sup> particles per second (for particles between 0.01 and 6 mm) is needed to explain the measured fluxes. Assuming a fragmentation power law for the size distribution, this corresponds to a dust production rate of $`5\times `$10<sup>7</sup> g s<sup>-1</sup>, in agreement with the theoretical estimates above (Landgraf et al. 2002). The study of hypervelocity micrometeoroid craters on lunar material and on the panels of the Long Duration Exposure Facility showed that Earth accretes about 3$`\times `$10<sup>7</sup> kg of interplanetary dust particles (IDPs) every year (Grün et al. 1985; Love & Brownlee 1993). Is the KB a significant source of these accreted IDPs? Kortenkamp & Dermott (1998, hereafter KD98) have calculated capture rates for IDPs of asteroidal and cometary origins. Based on these rates, and on the fact that the analysis of IDPs collected in the stratosphere shows a small diversity of chemical compositions (Flynn 1995), they argue that the sources of IDPs are very limited and lie mainly in the asteroid belt, with less than 25% having cometary origin. KB dust particles were, however, not considered in their study. The possibility that the KB may be a significant source of IDPs and the zodiacal cloud was first suggested by Liou, Zook & Dermott (1996, hereafter LZD96). They found that (1) about 20% of KB dust particles are able to reach the inner solar system and (2) these particles have small eccentricities and inclinations (similar to asteroidal grains) when they cross the orbit of the Earth, enhancing their chances of being captured and of surviving atmospheric entry. The significance of these results is that, as they explain, asteroids are certainly an important source of IDPs, but they can accrete material from only as far as $``$4 AU and it is not clear that organic material at such distances can survive the T-Tauri wind of the young Sun. KB dust grains, on the other hand, can bring in unaltered primitive material from the outer solar system, so they could potentially be a source of the earliest organic material that gave rise to life on Earth. Comets could also bring material from the outer solar system, but as LZD96 argue, their high eccentricities and inclinations cause the released dust particles to be in highly eccentric and inclined orbits. This translates into high encounter velocities with Earth ($`>`$10 km/s), making it difficult for any organic material to survive atmospheric entry. In their paper, however, they do not actually calculate capture rates and entry velocities for KB dust grains, giving only qualitative estimates. In this paper, we are going to follow numerically, from source to sink, the evolution of several hundred of dust particles from the KB under the combined effects of solar gravity, solar radiation pressure, Poynting-Robertson (P-R) and solar wind drag, and the gravitational forces of 7 planets (excluding Mercury and Pluto). The parent bodies of the dust particles are assumed to be distributed in orbits with semimajor axis between 35 and 50 AU, eccentricities such that the perihelion distances are between 35 and 50, and inclinations between 0 and 17, in approximate accord with current estimates of the orbital distribution of KBOs (Malhotra et al. 2000). We consider particles of diameter 3–115 $`\mu `$m (assuming $`\rho `$=1 g/cm<sup>3</sup>; equivalently 1–40 $`\mu `$m for $`\rho `$=2.7 g/cm<sup>3</sup>). The different particle sizes are referred to in terms of their $`\beta `$ value, which is the dimensionless ratio of the radiation pressure force and the gravitational force. For spherical grains and a solar type star, $`\beta `$=5.7 $`\times `$ 10<sup>-5</sup> Q<sub>pr</sub>/($`\rho `$b), where $`\rho `$ and b are the density and radius of the grain in cgs units (Burns, Lamy & Soter 1979). The quantity Q<sub>pr</sub> is the radiation pressure coefficient, a function of the physical properties of the grain and the wavelength of the incoming radiation; the value we use is an average integrated over the solar spectrum. The sinks of dust included in our numerical simulations are (1) ejection into unbound orbits, (2) accretion into the planets, and (3) orbital decay to less than 0.5 AU heliocentric distance. A detailed description of the models and the numerical algorithm used to integrate the equations of motion is given in Moro-Martín & Malhotra (2002). In $`\mathrm{\S }`$2, we present the radial density profiles derived from our models and the KB dust production rate from Landgraf et al. (2002); from these, we estimate the contribution of the KB dust to the zodiacal cloud. Assuming steady state, this leads us to an estimate of the total mass in the KB dust disk. We also describe how the effects of radiation forces and the planets’ perturbations change the particle size distributions. In $`\mathrm{\S }`$3, we address the question of the KB contribution to the collected IDPs on Earth by calculating geocentric encountering velocities and capture rates. In $`\mathrm{\S }`$4, we present the velocity field of the KB dust in the inner and outer solar system. In $`\mathrm{\S }`$5, we evaluate the magnitude of the Lorentz force, which is not included in our models. Finally, $`\mathrm{\S }`$6 summarizes our results. ## 2 Radial Distribution and Mass of the KB Dust Disk Based on our models and on estimates of KB dust production rates, we can calculate the number density of dust in the KB and its mass. The simulations yield radial profiles of the number density of dust for various particle sizes; the production rates are used to get the normalization of these profiles. For the production rates in the KB, we use the observationally based estimate by Landgraf et al. (2002) of 2$`\times `$10<sup>14</sup> particles per second of radius between 0.01 and 6 mm. For the size distribution, we use a fragmentation power law, $`n(b)\mathrm{𝑑𝑏}=n_\mathit{0}b^q\mathrm{𝑑𝑏}=n_\mathit{0}b^{\mathit{3.5}}\mathrm{𝑑𝑏}`$, where $`b`$ is the particle radius. (This corresponds to a generic grain mass distribution in collisional equilibrium; see, e.g., Dohnanyi 1969.) Then, assuming a bulk density $`\rho `$=1 g/cm<sup>3</sup>, we calculate the following dust production rates (in particles per second with the size bins in parentheses; the size corresponding to the particle’s diameter): 4.5$`\times `$10<sup>16</sup> (2.1–4.2 $`\mu `$m; $`\beta `$=0.4), 8.0$`\times `$10<sup>15</sup> (4.2–8.5 $`\mu `$m; $`\beta `$=0.2), 1.4$`\times `$10<sup>15</sup> (8.5–17 $`\mu `$m; $`\beta `$=0.1), 2.5$`\times `$10<sup>14</sup> (17–33.9 $`\mu `$m; $`\beta `$=0.05) and 4.9$`\times `$10<sup>12</sup> (81.3–162.6 $`\mu `$m; $`\beta `$=0.01). Because of the power law distribution, the size bins are chosen in such a way that they all have the same width in logarithmic scale; for each size bin, we have assigned a single $`\beta `$-value, as indicated (from Liou & Zook 1999). ### 2.1 Radial Distribution The radial profiles of the number density of particles within 1 AU of the ecliptic, based on our dynamical models and normalized using the dust production rates and the size bins listed above, are shown in Figures 1a and 1b. In our models, the parent bodies are assumed to be distributed in orbits with semimajor axis and perihelion distances between 35 and 50 AU. Upon release, a dust particle has the same position and velocity as its parent body, but its orbital elements are different as a result of radiation pressure. The latter effectively causes the particle to feel a Sun less massive by a factor (1-$`\beta `$). The larger the $`\beta `$, the more its orbit differs from its parent body’s. After a particle leaves its parent body, P-R drag and solar wind drag tend to circularize and decrease the semimajor axis of its orbit, forcing particles to slowly drift in toward the central star (Burns, Lamy & Soter 1979). Assuming that the dust particles are being produced constantly, this drif creates a dust disk of wide radial extent. Figure 1a shows the radial distribution of particles of five different values of $`\beta `$ in a fictitious KB disk unperturbed by planets. The radial distribution changes in the presence of planets. We have modeled the gravitational effects of seven of the planets, Venus through Neptune. Two effects play a major role in the quasi-steady state distribution of KB dust that we obtain in our models (see Fig. 1b): 1. Gravitational resonances.— The journey of the dust particles toward the central star is temporarily interrupted by the trapping of particles in Mean Motion Resonances (MMRs), mainly with the outermost planet, Neptune. The particles accumulate at certain semimajor axes, leading to the creation of structure in the disk; this explains the “bumps” that appear between 35 and 50 AU. The structure is more prominent for the smaller $`\beta `$-values because the resonance trapping is more efficient when the drag force is small. 2. Gravitational scattering.— Massive planets scatter and eject dust particles out of the planetary system, an effect that is independent of $`\beta `$. Scattering is responsible for the most striking difference between Figures 1a and 1b: for particles larger than the blow-out size ($`\beta `$$``$0.5), the scattering of dust by the giant planets is able to extend the disk beyond the boundaries set by radiation effects alone. This has important consequences on the size distribution that will be explored below. In Figure 1c, we show the radial profile of the total number density of particles with radius between 1.4 and 10 $`\mu `$m, compared to the Voyager 1 number density estimate, inferred from dust impact rates of 1.4–10 $`\mu `$m particles from 30 to 51 AU (Gurnett et al. 1997; Jewiit & Luu 2000). The radial profiles are obtained as follows: First, for each choice of particle bulk density $`\rho `$, we assign each $`\beta `$-value to a particle size bin (the size being the particle’s diameter). For $`\rho =1`$ g/cm<sup>3</sup>, we define the following size bins: 2.8–4.2 $`\mu `$m ($`\beta `$=0.4); 4.2–8.5 $`\mu `$m ($`\beta `$=0.2), 8.5–17 $`\mu `$m ($`\beta `$=0.1) and 17–20 $`\mu `$m ($`\beta `$=0.05). For $`\rho =2.7`$ g/cm<sup>3</sup>, the size bins are: 2.8–5.3 $`\mu `$m ($`\beta `$=0.1), 5.3–13.3 $`\mu `$m ($`\beta `$=0.05) and 13.3–20 $`\mu `$m ($`\beta `$=0.01). Next, we normalize the radial profile obtained from our numerical models for each of the values of $`\beta `$ (i.e., for the corresponding size bins) using the dust production rates from Landgraf et al. (2002) and assuming the power law size distribution stated above (q=3.5). As a final step, we add the contribution from all the size bins to obtain the total number density radial distribution shown in Figure 1c. The differences between the profiles for the two choices of $`\rho `$ arise from the fact that the bulk density affects the correspondence between the $`\beta `$-value and the particle size, and the size in turn affects the estimate of the dust production rate because of the assumed power law in the size distribution. When comparing the modeled radial profiles with the Voyager 1 estimate, one should keep in mind that there are uncertainties in the dust production rates and in the index of the power law (both of which determine the normalization factors of the models), as well as in the Voyager 1 number density estimate (which is based on a few impact events and also has some uncertainty in the sizes of the particles detected). Nevertheless, the two modeled radial profiles are in good agreement with the Voyager 1 observations. We cannot favor silicate over icy composition for KB particles based on this comparison. ### 2.2 Size Distribution Radiation forces and planetary perturbations change the size distribution of dust particles, as the particles spread out from their site of production at rates that are dependent on their size. Figure 2 shows these effects in plots of the cumulative size distribution at various heliocentric distances throughout the KB dust disk. (The cumulative size distribution is calculated by integrating the differential size distribution obtained from our models in the size bins described above for $`\rho `$=1 g/cm<sup>3</sup>). As we mentioned, the initial differential size distribution at the time of dust production is assumed to be a power law with q=3.5; it is represented in Figure 2 as the thick line of slope -2.5, with the distance between the squares indicating our particle size binning “resolution”. The other lines represent the cumulative size distribution obtained in our models at five different heliocentric distances: 5, 21, 41, 81 and 141 AU, as indicated in the figure. Figure 2a shows results for a fictitious KB disk unperturbed by planets, while Figure 2b shows results for the KB dust disk perturbed by the seven planets, Venus through Neptune. The main features are the following: 1. Radiation forces alone change the differential size distribution, from the original power law with q = 3.5, to another power law of smaller index (see Fig. 2a). This is due to the fact that radiation pressure “kicks out” the smaller particles preferentially and P-R drag spreads out the smaller particles faster than the bigger ones. 2. This shallower power law (with slope of $``$-1.5, corresponding to a differential power law index q$``$2.5) is maintained constant throughout the disk at distances smaller than the aphelion of the parent bodies (parallel dotted, dashed and solid lines in Fig. 2a). At larger distances, however, we start to encounter the disk boundaries set by radiation pressure, which depend on the particle sizes. This explains the steeper size distributions found at 81 and 141 AU (only the smaller particles reach those larger distances). 3. In the presence of planets, the size distribution changes greatly at distances larger than the aphelion of the parent bodies: compare the shallower slopes obtained at 141 AU and 81 AU in Figure 2b, with the steep slopes at the same distances in Figure 2a. Unlike radiation pressure, gravitational scattering by the giant planets can send larger particles to these larger distances, effectively spreading all the dust widely. As the figure shows, the dust distribution is no longer described by a power law with a single index. 4. The trapping of particles in MMRs with Neptune (between 35 and 50 AU), and the fact that large particles are more easily trapped, explains why the slope of the size distribution becomes more shallow at 41 AU (slope about -1.5, q$``$2.5) than at 5 and 21 AU (slope about -1.9, q$``$2.9) (compare solid and dashed lines with dotted line in Fig. 2b). Although some of these effects are minor, the large change in the size distribution described in point 3 is very significant. It is clear that the detection of an exoplanetary dust disk of wide radial extent (a hundred to thousands of AU) does not necessarily imply the presence of dust-producing planetesimals at such large distances: gravitational scattering by giant planets can spread the dust to distances much larger than the aphelion of the parent bodies. The obvious question is whether this effect could be used to unambiguously infer the presence of giant planets. We plan to address this question in the future by studying the effect that the change in the particle size distribution has on the disk’s spectral energy distribution. ### 2.3 Total Dust Mass From our models, we estimate the total mass of the KB dust disk to be, $`m_{KBdust}`$1.2$`\times `$10<sup>-11</sup> M for particles with diameter 2.4–160 $`\mu `$m (assuming $`\rho `$=1 g/cm<sup>3</sup>), or $`m_{KBdust}`$4.2$`\times `$10<sup>-11</sup> M for particle with diameter 0.8–150 $`\mu `$m (for $`\rho `$=2.7 g/cm<sup>3</sup>). Using COBE observations at 140 and 240 $`\mu `$m, Backman, Dasgupta & Stencel (1995) set an upper limit for the total mass of dust in the KB of $``$ 3$`\times `$10<sup>-10</sup> M. Jewitt & Luu (2000) calculated that the mass in particles with radius 1.4–10 $`\mu `$m to be $``$4$`\times `$10<sup>-14</sup> M, based on a simple estimate using the Voyager 1 number density; the volume of an annulus with 30 AU inner radius, 50 AU outer radius, and 10 AU thickness; and the assumption of an average grain mass of 2$`\times `$10<sup>-14</sup> kg. For this size range, and using the same size bins used for Figure 1c, our models predict a mass of 5.2$`\times `$10<sup>-13</sup> M (for $`\rho `$=1.0 g/cm<sup>3</sup>), or 5.5$`\times `$10<sup>-12</sup> M (for $`\rho `$=2.7 g/cm<sup>3</sup>). The uncertainties in the derived $`m_{KBdust}`$ come not only from the dust production rates, but also from the fact that we are extrapolating the results from only five $`\beta `$-values to a wide range of particle sizes. To estimate $`m_{KBdust}`$, we do the following: (1) We count the number of particles present in our five “steady state” models, each corresponding to a different $`\beta `$. The models assume an artificial dust production rate of 100 particles every 1000 years (see Moro-Martín & Malhotra 2002 for more details). (2) We multiply the number of particles by the ratio of the dust production rates derived from Landgraf et al. (2002; in the size bins corresponding to the values of $`\beta `$ under consideration), to our artificial dust production rate. This gives us the total number of particles in each of the five size bins. (3) To convert this number into mass, one must multiply by the particle mass. The particle mass that we attribute to each size bin is calculated using the fragmentation power law, so that the small particles have more weight because they are more abundant. (If we were to use the mass of the particle that lies in the middle of the bin \[corresponding to the modeled $`\beta `$\], our total dust masses would be about 4.5 times larger). (4) Finally, we add together the masses from the five different size bins. This results in the values of $`m_{KBdust}`$ quoted above. ## 3 Is the KB a Significant Source of IDPs? We have calculated Earth’s capture rates and entry velocities for KB dust grains based upon our numerical models, and adopting the procedure of KD98. We find that (1) KB dust grains have higher eccentricities when crossing the orbit of the Earth than those found by LZD96 and (2) their encounter velocities and capture rates are more similar to dust grains of cometary origin than to asteroidal origin; this is contrary to the results of LZD96. We define a particle to be Earth crossing if its orbit overlaps that of Earth, that is, $`q<`$$`R`$$`<Q`$, where $`q`$=$`a`$(1-$`e`$) is the perihelion of the particle, $`Q`$=$`a`$(1+$`e`$) is its aphelion and R is the heliocentric distance of Earth, 0.9833 AU$`<`$R$`<`$ 1.0167 AU. The encounter velocity $`v_\mathit{0}`$ between the Earth and a particle on a crossing orbit was calculated following Kessler (1981). The effective capture cross-section $`\sigma _c`$ is given by $`\sigma _c`$=$`\sigma _{\mathrm{}}`$(1+$`v_{e}^{}{}_{}{}^{\mathit{2}}`$/$`v_{\mathit{0}}^{}{}_{}{}^{\mathit{2}}`$), where $`v_e`$ is the escape velocity from the Earth (at an altitude of 100 km, $`v_e`$=11.1 km s<sup>-1</sup>) and $`\sigma _{\mathrm{}}`$ is Earth’s geometric cross section. The average spatial density at heliocentric distance R and ecliptic latitude $`l`$ of a population of dust particles with orbital elements $`a`$, $`e`$ and $`i`$ is given by $$S(R,l)=\frac{1}{2\pi ^3Ra[(\text{sin}^2i\text{sin}^2l)(Rq)(QR)]^{1/2}}.$$ (1) The fraction of this population captured by the Earth at position (R,l) per unit time is p=$`v_\mathit{0}`$$`\sigma _c`$S(R,l). Following KD98, for each of the Earth-crossing particles in our models we have calculated S(R,l) at 360 positions along Earth’s orbit, with R and $`l`$ uniformly distributed in the range 0.9833 AU$`<`$R$`<`$ 1.0167 AU and -0<sup>o</sup>.00035$`<l<`$ 0<sup>o</sup>.00035. Table 1 shows the results after averaging over these 360 positions and over the whole population of Earth-crossing particles. For comparison, the results from KD98 for asteroidal and cometary dust (with $`\beta `$=0.0469) are also included. As Figure 3 shows, we find high eccentricities for KB dust grains, similar indeed to cometary dust and not to asteroidal dust, which implies a low spatial density and high encounter velocity, and, therefore, a low capture rate (see Table 1). The asteroidal dust particles, on the other hand, have lower eccentricities and inclinations, which translates into a higher capture rate. The discrepancies with LZD96 probably arise from the different criterion used to identify particles on Earth-crossing orbits; LZD96’s criterion, $`a<`$1, which is less precise than the one we adopt here, has a strong bias toward low-eccentricity orbits. In order to estimate the relative contributions of various sources to the IDPs captured at Earth from the relative capture rates in Table 1 (which depends only on the orbital elements of the population of Earth-crossing particles), we need to know the relative contribution of each source to the number density of particles on Earth-crossing orbits. This problem has yet to be solved because the actual dust production rates from asteroids, comets and KBOs are highly uncertain and very model dependent. Since cometary and KB dust grains have similar capture rates, we can extend the results of KD98 to predict that cometary plus KB dust can represent more than half of the IDPs captured by Earth only if comets and KBOs together supply $``$95% of the Earth-crossing particles. Based on modeling of the IRAS dust bands, KD98 estimated that (5–25)% of the Earth-crossing particles originate in the asteroid families Eos, Themis and Koronis. Using Dermott et al.’s (1994) 1:3 ratio of dust produced by asteroid families to that produced by all asteroids, we then have that all asteroids contribute (15–75)%, leaving the rest for comets and KBOs. In the extreme case that as much as 85% has cometary or KBO origin, this suggests that because of the lower capture rates of these highly eccentric grains, only 25% of the collected IDPs will be supplied by comets and KBOs. Interestingly, Brownlee, Joswiak & Love (1994) concluded that, based upon the maximum temperature reached during atmospheric entry from the study of helium release, about 20% of IDPs $`<`$ 10 $`\mu `$m have entry velocities typical of cometary IDPs. This is in agreement with the above estimate. In the other extreme case, where 25% of IDPs have cometary or KBO origin, our models, together with KD98’s results, suggest that they will represent only about 2% of the collected IDPs. Our conclusion from this exercise is that the KB can certainly be a source of IDPs but it is not as important as predicted by LZD96. The estimates above are for the relative contributions from the different sources to the collected IDPs and depend on their relative contribution to the number density of particles on Earth-crossing orbits. We can also calculate the absolute contribution from the KB by using Landgraf et al. (2002) dust production rates and the capture rates in Table 1. Our models, together with the dust production rates, yield the number of particles on Earth-crossing orbits. The capture rates are the fraction of this population that is captured by Earth every 10<sup>9</sup> yr. The multiplication of these two numbers leads to the following results: 1.2$`\times `$10<sup>5</sup> kg yr<sup>-1</sup> (2.4–160 $`\mu `$m, $`\rho `$=1 g cm<sup>-3</sup>), or 4.1$`\times `$10<sup>5</sup> kg yr<sup>-1</sup> (0.8–150 $`\mu `$m, $`\rho `$=2.7 g cm<sup>-3</sup>). These numbers should be compared to the total mass influx of 3$`\times `$10<sup>7</sup> kg yr<sup>-1</sup> inferred from the microcraters on the Long Duration Exposure Facility (Love & Brownlee 1993). The microcraters correspond to particles of radius between 2.5 and 250 $`\mu `$m, and show a peak and a cutoff in the particle size distribution near 100 $`\mu `$m. The accreted KB dust mass represents between 0.4% (assuming $`\rho `$=1 g cm<sup>-3</sup>) and 1.4% (assuming $`\rho `$=2.7 g cm<sup>-3</sup>) of this total mass influx. The same uncertainties in the dust mass estimates mentioned in $`\mathrm{\S }`$2.3 apply here —namely, if we were to use the mass of the particle that lies in the middle of each bin instead of weighting the mass using the power law, the values would be 4.5 times larger. Also, because most of the mass is contained in the large particle sizes, these results depend on the maximum particle radius chosen. The conclusion, however, is clear: if Landgraf et al. (2002) KB dust production rates are correct, then the KB presently provides only a few percent of the collected IDPs. The delivery rate of KB dust to Earth’s vicinity is calculated using the dust production rates in $`\mathrm{\S }`$2 and the percentage of particles that is able to drift all the way into the Sun, which, as seen in Table 2, is also a function of $`\beta `$. The delivery rates are (expressed in particles per second, followed by the bin sizes in parentheses) 4.9$`\times `$10<sup>15</sup> (2.1–4.2 $`\mu `$m), 1.2$`\times `$10<sup>15</sup> (4.2–8.5 $`\mu `$m), 2.9$`\times `$10<sup>14</sup> (8.5–17 $`\mu `$m), 4.7$`\times `$10<sup>13</sup> (17–33.9 $`\mu `$m) and 5.4$`\times `$10<sup>11</sup> (81.3–162.6 $`\mu `$m). One should keep in mind, however, that these estimates, and the ones in $`\mathrm{\S }\mathrm{\S }`$ 2 and 4, are rather model-dependent: the Landgraf et al dust production rate estimate makes assumptions about the KB parent bodies’ orbits that are significantly different from the observed distribution; they also assume that there is no source of dust in the 10–30 AU region, and we have neglected the destruction of dust grains due to interstellar and mutual collisions. ## 4 Velocities of KB Dust Grains A study of the velocity field of KB dust is useful for predicting the flux of particles colliding with a spacecraft exploring the outer solar system (e.g. New Horizons and Interstellar Probe).<sup>1</sup><sup>1</sup>1http://pluto.jhuapl.edu/ and http://interstellar.jpl.nasa.gov/ This is of interest for planning dust detectors or dust analyzers, as well as for estimating the potential hazard posed by dust collisions to fast-moving spacecraft. In order to provide some general estimates, we have used our models to calculate the noncircular velocity of the KB dust in the ecliptic: for each particle, the instantaneous circular velocity at that distance has been calculated and has been subtracted from its actual velocity. The resulting magnitude of noncircular velocities in the ecliptic presented in Figure 4 corresponds to the average values of the particles that lie in square cells of 1 AU in size. We find no significant azimuthal structure, except for the following: between 25 and 35 AU, the non-circular velocities show a small systematic azimuthal variation at the level of 10–20%, with a maximum at Neptune’s position, which may be due to the fact that the particles trapped in MMRs tend to avoid the planet in the resonance. Figure 5 shows the radial profile of the ratio between the non-ircular and the circular velocity averaged over azimuth. The increase of the fractional noncircular velocity for heliocentric distances $`r`$50 AU is due to the fact that only particles of large eccentricities are to be found at distances beyond the parent bodies. The noncircular velocities tend to be higher for smaller particles (larger $`\beta `$), as expected from their larger eccentricities upon release. ## 5 Other Physical Processes Our models do not include the effect of magnetic fields on charged dust grains and the dust grain destruction processes (such as sublimation, sputtering and collisions). Below, we briefly comment on how this may affect the results presented here, but a comprehensive evaluation of these processes is beyond the scope of this paper. ### 5.1 Effect of Heliospheric Magnetic Fields Dust grains are generally electrically charged, as a results of the ejection of photoelectrons and the accretion of ions and electrons. Inside the heliosphere, $``$ 150 AU, the grains are therefore subject to the Lorentz force exerted by the interplanetary magnetic field, while outside the heliosphere the interstellar magnetic field dominates. The effects of solar wind magnetic forces on charged dust grains have been discussed previously (Parker 1964; Consolmagno 1979; Morfill & Grün 1979; Mukai 1985; Gustafson 1994; Fahr et al. 1995; Grün & Svestka 1996). Here we summarize the lines of argument that lead us to conclude that the omission of Lorentz forces in our modeling is not a significant limitation of our results. The interplanetary magnetic field is known to have a complex structure and time behavior. The dipole component changes polarity every 11 years, with the 22 year solar cycle. Moreover, near the ecliptic these sign reversals take place more rapidly because of the presence of the heliospheric current sheet, the extension of the Sun’s magnetic equator into interplanetary space, separating regions of opposite polarity. At solar minimum, the current sheet extends from approximately $`25^{}`$ to $`25^{}`$ from the solar equator. Particles within this latitude range cross the current sheet at least twice every solar rotation ($``$27 days), or four or even six times if the current sheet is wrapped because of higher order terms in the magnetic field (Balogh 1996). At higher ecliptic latitudes, the particles cross the current sheet at least twice as they orbit the Sun. Therefore, the time-averaged effect of the Lorentz force will tend to vanish within a particle’s orbital period, because the sign reversals are significantly faster than the orbital period of most KB particles. (We note that 80–90% of the KB dust grains are ejected by the giant planets \[see Table 2\], and therefore their orbital periods during their lifetimes are generally much larger than the 11.8 year period of the innermost giant planet, Jupiter. However, we cannot rule out resonant effects for charged grains that remain in the vicinity of Jupiter and Saturn for extended periods of time, as they may be subject to Lorentz forces of period comparable to their orbital periods.) Parker (1964) was the first to study the effect of this fluctuating interplanetary field on dust grains on non-inclined, circular orbits. Because the dominant component of the field is perpendicular to the radial solar wind vector, with a magnitude $`3\times 10^5/r`$(AU) gauss for heliocentric distances r exceeding a few AU (Parker 1963, p. 138), he concluded that the Lorentz force will scatter the grains out of the ecliptic plane, by perturbing the particle’s inclinations while keeping the energy of the orbit unchanged. At the distance of the Earth, the scattering would be important only for grains $``$1 $`\mu `$m, for which the inclinations change significantly before P-R drag sweeps them into the Sun. More recently, Fahr et al. (1995) estimated that the inclination change causes a negligible evolutionary effect on zodiacal dust particles $``$10 $`\mu `$m. They found that for particles with inclinations $`i`$$``$15, where the bulk of the dust particles considered in this paper are, this effect is completely negligible compared to P-R migration rates because of the stochastic character of the electromagnetic force near the current sheet; for $`i`$$`>`$15 and circular or quasi-circular orbits, the Lorentz force exactly cancels out when integrated over a complete orbit, whereas for more eccentric orbits, the orbit-averaged change in inclination turns out to be very small because the Lorentz force reverses every 11 years with the solar cycle. But as Parker (1964) pointed out, in reality, the interplanetary field also fluctuates in the direction perpendicular to the ecliptic. These fluctuations cause a random walk in the semimajor axis of the particles. Over a period of time $`\mathrm{\Delta }t`$, the P-R effect will dominate over Lorentz scattering provided that $`\mathrm{\Delta }a_{PR}\mathrm{\Delta }a^2_L^{1/2}`$. Using Consolmagno’s (1979) derivation for $`\mathrm{\Delta }a^2_L^{1/2}`$ in a circular orbit, Jokipii & Coleman’s (1968) estimates for the fluctuating perpendicular component of the magnetic field (based on measurements by Mariner 4), Burns et al.’s (1979) expression for $`\mathrm{\Delta }a_{PR}`$, and adopting $`q=bV/300`$ esu for the particle’s electric charge, we can write the condition above as $`\frac{bQ_{pr}}{V}0.64\left(\frac{a^3}{\mathrm{\Delta }t}\right)^{1/2}`$, where $`b`$ is the particle’s radius in $`\mu `$m, V is its electrical potential in volts, $`a`$ is in AU, and $`\mathrm{\Delta }t`$ is in years. The dependence on the time $`\mathrm{\Delta }t`$ arises from the fact that the P-R effect causes a systematic drift in a that is proportional to $`\mathrm{\Delta }t`$, while the fluctuating Lorentz force causes a diffusion in a that is proportional to $`(\mathrm{\Delta }t)^{1/2}`$. Scaling the comparison time $`\mathrm{\Delta }t`$ by the orbital period of the particle, that is, $`\mathrm{\Delta }t=(ka)^{3/2}`$, where $`k`$ is a numerical factor, we find that the P-R effect will dominate Lorentz scattering for particle sizes $`b3.2\left(\frac{a/1\text{AU}}{k}\right)^{3/4}\left(\frac{V}{5\text{volt}}\right)\left(\frac{1}{Q_{pr}}\right)\mu \text{m}`$. Thus, for particles of radius larger than a few microns, the systematic P-R drift will exceed the random Lorentz scattering on timescales from a few orbital periods in the inner solar system to a few tens of orbital periods in the outer solar system. Over the characteristic P-R drift timescale, $`(a/\dot{a})_{PR}`$, Lorentz scattering is negligible for the particle sizes and heliocentric distances in our models. We therefore consider that neglecting the Lorentz force does not constitute a major limitation of this work. ### 5.2 Sublimation Silicate grains can survive sublimation to distances less than $`0.5`$ AU, whereas pure water-ice grains are unlikely to survive interior to $`4`$ AU (see estimates in $`\mathrm{\S }`$6.2 of Moro-Martín & Malhotra 2002). For example, the staying time for a grain between 1 AU and 2 AU from the Sun on a orbit with $`a`$=10 AU and $`e`$=0.9 (i.e. q=1AU and Q=19AU) is $``$10<sup>7</sup> s, while the lifetime of an icy grain with a radius of 163 $`\mu `$m is $``$10<sup>3</sup> s at 2 AU from the Sun (Mukai 1986; Gustafson 1994). Therefore, the icy grain cannot survive near the Earth due to quick sublimation. KB grains are likely a mixture of silicates and ices. While the ice fraction will sublimate quickly, the silicate remnant will likely survive to sub-Earth perihelion distances. Qualitatively, and for the size ranges considered in this paper, we expect that the rapid loss of the ice component will cause the grain’s orbit to become more eccentric, as a result of the increased magnitude of radiation pressure on smaller grain sizes. Thus, our dynamical models would underestimate the eccentricities of KB grains on Earth-crossing orbits. (However, for smaller grains of radii less than a few tenths of a micron, the effect would be the opposite because $`\beta `$ decreases as the grain’s radius decreases.) Furthermore, taking into account the sublimation of the icy fraction, our conclusion from $`\mathrm{\S }`$3, that $`1\%`$ of silicate IDPs may be from the KB, becomes an upper limit. The overall conclusion is still the same: most of the captured IDPs do not come from the KB. ### 5.3 Sputtering Sputtering by solar wind particles may cause mass loss and erosion of dust grains, as well as chemical alteration of their surfaces. The erosion rate is quite uncertain in existing literature. Most estimates are based on the analysis of Apollo samples of lunar soils and related computer simulations and bombardment experiments. Some of these estimates are as follows: McDonnell & Flavill (1974) and McDonnell et al. (1977), estimated an erosion rate of 0.043 $`\AA `$ yr<sup>-1</sup> and 0.43 $`\AA `$ yr<sup>-1</sup>, respectively, on the basis of He<sup>+</sup> bombardment experiments. A few years later, Flavill et al. (1980) estimated 0.025–0.045 $`\AA `$yr<sup>-1</sup>, while Kerridge (1991) estimated 0.0024 $`\AA `$ yr<sup>-1</sup> based on analysis of Ar<sup>36</sup> retention efficiency for solar wind implantation and its measures in a lunar sputtered surface. In another independent study, Johnson & Baragiola (1991) estimated erosion rates of 0.1–0.2, 0.01–0.03 and 0.002–0.003 $`\AA `$ yr<sup>-1</sup>, where the two lower estimates take into account the decrease of sputtering efficiency due to the sticking of sputtered material to neighboring grains and to micrometeorite vapor deposition, respectively. Evidently, the estimated erosion rates differ by up to a factor of 200 in these studies. Most recently, Mukai et al. (2001) suggest a rate of 0.1–0.2 $`\AA `$ yr<sup>-1</sup>. Adopting an erosion rate of 0.2 $`\AA `$ yr<sup>-1</sup> at 1 AU, and taking into account that it scales with heliocentric distance roughly as r<sup>-2</sup>, we can estimate the mass loss experienced by our modeled KB dust grains. Our dynamical studies of KB dust show that most of the particles spend most of their time at $`a`$$`>`$20 AU, and that their typical lifetime is $``$10<sup>7</sup> yr (Moro-Martín & Malhotra 2002 Figs. 3 and 10). Consider a typical particle that spends 10<sup>7</sup> yr at 20 AU from the Sun. The fraction of mass loss is $``$ 50% for a 3 $`\mu `$m particle, and it scales as b<sup>-1</sup>, where $`b`$ is the particle radius. (This is likely an upper limit because the particles usually get trapped in exterior MMRs with Neptune at $`a`$$`>`$30AU.) Of course, one would need to take into account that as the particles drift in as a result of P-R drag, their erosion rate increases because of increased solar wind flux at smaller heliocentric distance. Our dynamical studies show that typically particles spend less than $``$10<sup>6</sup> yr inside 20 AU. We estimate that a 3 $`\mu `$m grain will be almost completely destroyed before reaching the inner solar system, while a 10 $`\mu `$m grain will suffer little erosion. If the the erosion rate is 100 times smaller than our adopted value (and within the present uncertainties), the mass loss would be negligible in both cases. We may therefore conservatively conclude that grains $`>`$10 $`\mu `$m do not suffer significant erosion due to corpuscular sputtering. Sputtering-induced chemical alteration of dust grain surfaces may also reduce the mass loss. Corpuscular sputtering preferentially depletes the surface regions of volatiles, but also causes implantation of ions that can change the chemistry of the grain surface by producing mixing and molecular bonding between layers of dissimilar materials. This may explain why IDPs, thought to be Van der Waals-bonded aggregates, can lose icy mantles and remain sufficiently stable to survive atmospheric entry. A blackened, sputter-resistant, highly carbonized and refractory surface layer can be created from organic and volatile mantles (Johnson & Lanzerotti 1986, Johnson 1990, Mukai et al. 2001). Once this layer is formed, the efficiency of erosion by corpuscular sputtering will be reduced. Our conclusions above are consistent with the findings of Mukai & Schwehm (1981) and Johnson (1990) who conclude that at the distances at which sputtering is important, the erosion is relatively small under present solar wind conditions but chemical alterations may be significant. ### 5.4 Collisions In the optically thin limit of interest in the present work, mutual collisions of dust grains are not significant, but grain destruction due to collisions with interstellar grains may be (Liou & Zook, 1999). In Moro-Martín & Malhotra (2002; $`\mathrm{\S }`$6.1), we compared the collisional lifetimes estimated by LZD96 to the dynamical lifetimes derived from our models. We concluded that collisions with interstellar grains are likely to be important for KB dust particles with diameters from 6 to 50 $`\mu `$m: smaller particles survive because they drift in fast, and larger particles survive because they are not destroyed by a single impact. Interstellar grain collisions therefore may affect the particle size distributions presented in $`\mathrm{\S }`$2.2. It would be useful to address this in detail in a future study. 1. We note here that one of our long-term goals, as part of the SIRTF FEPS Legacy project (principal investigator M. Meyer),<sup>2</sup><sup>2</sup>2http://feps.as.arizona.edu is to study the effect of planets and radiation on the particle size distribution in exo-planetary systems. Considering that there are large uncertainties in the solar wind corpuscular sputtering effects, as well as the interstellar grain flux and size distribution for our own solar system, we think it is not well-justified to introduce in our numerical models the effects of sputtering and collisions for systems where the interstellar dust environment would likely be even less well known. ## 6 Summary and Conclusions 1. We have estimated the radial distribution of KB dust from our dynamical models and the KB dust production rate estimates from Landgraf et al. (2002). (We neglect dust physical destruction processes.) We find that the presence of planets has a very important effect on the distribution of dust: for particles larger than the blow-out size ($`\beta 0.5`$), the gravitational scattering of dust by the giant planets is able to extend the disk beyond the boundaries set by radiation effects alone. We also find that it has important consequences for the dust size-frequency distribution (see below). 2. The observation of dust disks of wide radial extent, a hundred to thousands of AU, does not necessarily imply the presence of dust-producing planetesimals at such large distances, because the gravitational scattering by giant planets at much smaller semimajor axes can cause the dust to spread to distances much larger than the aphelion of the dust parent bodies. 3. Radiation forces alone change the differential size distribution from the (assumed) initial power law of index $`q=3.5`$ at production, to a shallower power law with $`q2.5`$, valid at distances smaller than the aphelion of the parent bodies. No large particles are found at larger distances, and consequently the size distribution there is very steep. However, when we account for planetary perturbations, the size distribution changes greatly at these large distances. Overall, we conclude that the combination of radiation forces and planetary perturbations causes the dust disk to spread out and the dust size frequency distribution to flatten (Figs. 1 and 2). In a future study, we plan to investigate the potential of the latter effect for the detection of planets in debris disks. 4. We estimate the total mass of the KB dust disk to be $`m_{\mathrm{KB}\mathrm{dust}}1.2\times 10^{11}`$ M (2.4–160 $`\mu `$m, $`\rho `$=1 g cm<sup>-3</sup>), or $`m_{\mathrm{KB}\mathrm{dust}}4.2\times 10^{11}`$ M (0.8–150 $`\mu `$m, $`\rho `$=2.7 g cm<sup>-3</sup>). These estimates are consistent with other KB dust mass estimates found in the literature. 5. We find in our dynamical models that KB dust grains near Earth have high eccentricities and inclinations similar to those of cometary grains and not asteroidal grains (Fig. 3). (Sublimation of the volatile fraction of these grains in the inner solar system is likely to increase their eccentricities further.) As a consequence, they have encounter velocities and capture rates similar to cometary dust values; this is contrary to previous results (Liou et al. 1996). 6. We estimate, following Kortenkamp & Dermott (1998), that at most 25% of IDPs captured by Earth have cometary or KB origin. Furthermore, using Landgraf et al.’s (2002) estimates of KB dust production rates, we find that the KB presently provides no more than a few percent of the collected IDPs. 7. We have present the velocity field of KB dust grains in the inner and outer solar system (Figs. 4 and 5). This is potentially useful for planning dust detectors on future spacecraft missions, as well as for estimating the hazard to space probes in the outer solar system. 8. We estimate that the Lorentz forces due to the interplanetary magnetic field within the heliosphere are likely negligible for the particle sizes considered in this paper. Mainly as a result of the rapid reversals in magnetic field polarity with the solar cycle, and the wrapped structure of the heliospheric current sheet, the effect of the Lorentz force will tend to average out within a particle’s orbit. 9. Some physical destruction processes on KB dust grains may affect their dynamical evolution significantly, and detailed analysis in warranted in future studies. We estimate that the effect of rapid sublimation of the volatile component of KB dust grains is to increase their Earth encounter velocities and to reduce their relative abundance among captured IDPs. The effects of sputtering by the solar wind are insignificant for grain sizes exceeding $`10`$$`\mu `$m. Collisional destruction by interstellar grains likely modifies the size frequency distribution further, beyond the effects considered in our dynamical models. Acknowledgments We thank Hal Levison for providing the SKEEL computer code and R. Jokipii and the anonymous referee for helpful discussions and comments. A. M.-M. thanks the SIRTF Science Center and IPAC for providing access to their facilities during the completion of this work. A. M.-M. is supported by NASA contract 1224768 administered by JPL. R. M. is supported by NASA grants NAG5-10343 and NAG5-11661.
warning/0506/hep-ph0506261.html
ar5iv
text
# Mixed top-bottom squark production at the LHC ## I Introduction Spacetime Supersymmetry (SUSY) has become one of the leading candidate extensions to the Standard Model (SM) of particle physics SUSY . Besides the aesthetic appeal of maximally extending the Poincaré group Haag:1974qh , the Minimal Supersymmetric Standard Model (MSSM) also provides a natural weakly-interacting dark matter candidate, a solution to the Higgs sector naturalness problem, and likely unification of gauge couplings near the Planck scale, among other features. If SUSY exists, it must be a broken symmetry at low energy, as we do not observe the opposite-spin-statistics partners of the SM field content which would necessarily exist. As a result, the squarks, sleptons, charginos and neutralinos of the MSSM must be fairly massive in comparison to their SM counterparts. Previous high-energy physics experiments such as LEP and Run I of the Tevatron have put stringent bounds on some of the sparticle masses. It will fall to the CERN LHC proton-proton collider to perform a conclusive SUSY search, although the real physics would only begin were a potential SUSY discovery to be made. If LHC finds candidate SUSY particles, a great deal of work would be required to prove or disprove the hypothesis that the new particle content belongs to the SUSY partner spectrum of the SM content, and if confirmed to perform a sufficient number of measurements to determine the parameters of the SUSY Lagrangian Bechtle:2004pc ; Lafaye:2004cn ; Gjelsten:2005aw . For the LHC the prospects of undertaking both tasks is uncertain, since many of the weakly-interacting spartners might go unobserved, and many of the colored spartners would have overlapping signatures; disentangling these would be a daunting prospect. The LHC will also be hard pressed to measure sparticle masses: for some, only mass differences may be accessible Hinchliffe:1996iu ; Bachacou:1999zb ; Allanach:2000kt ; Gjelsten:2005aw . Performing such tasks as measuring the spins of new state, or confirming that their couplings are indeed identical to the gauge and Yukawa couplings of the SM, is largely unexplored territory phenomenologically. Ideally, a future linear collider (ILC) will be constructed which could address many of these precision measurement problems ILC , and there is considerable effort to understand the synergy that would exist between LHC and ILC Weiglein:2004hn . However, there is also now the planned luminosity upgrade to LHC, the SLHC Gianotti:2002xx . While it certainly won’t be able to bring about a precision-era of physics and guarantee disentanglement of whatever new physics is found at LHC, our goal here is to explore just how far LHC and SLHC could push the envelope in measurements of the new physics. SUSY phenomenology at LHC has so far focused mainly on the dominant $`22`$ processes, which for colored sparticles is QCD production of gluino pairs, same-flavor squark pairs, and mixed gluino-squark pairs. These production channels probe only the SUSY-QCD vertices. While the sparticle cascade decays will necessarily involve weak-interaction vertices, the actual measurement is the number of events resulting from the production cross section times branching ratio (BR) to a given final state. This alone is not enough information to separate the different vertices. <sup>1</sup><sup>1</sup>1In principle, the QCD vertex could be measured by observing all possible sparticle decays, so that the total rate observed is $`100\%`$ of the BR. Analogous to the case of single-top production in the SM Chakraborty:2003iw , observing mixed-flavor squark production, which occurs via the weak vertex $`W`$-$`\stackrel{~}{q}_L`$-$`\stackrel{~}{q}_L^{}`$ and semi-weak vertex $`W`$-$`g`$-$`\stackrel{~}{q}_L`$-$`\stackrel{~}{q}_L^{}`$, would add additional information about the sparticle interaction vertices. Note that the weak interaction, being left-handed, involves only the left-handed squark partners of the quarks. In fact, assuming that the various sparticle BRs were determined in QCD pair production, mixed-flavor production would then provide absolute measurements of the weak vertex couplings. Weak production processes are naïvely approximately two orders of magnitude smaller than QCD production processes - although this also depends on the relative mass spectrum with respect to which parton luminosity dominates for that regime of $`x`$. This would present a problem for identifying this mixed-flavor signal. There is some hope, however. In maximally-unifying scenarios, a.k.a. “GUT-inspired”, over much of parameter space the colorless sparticles would be lighter than the colored sparticles. The cascade decay chains of squarks would typically therefore proceed ultimately through the lightest slepton (almost always a stau in popular unifying scenarios) or lightest chargino, in either case giving a final-state lepton <sup>2</sup><sup>2</sup>2A third possibility exists, where the cascade proceeds ultimately through an on-shell $`Z`$ boson plus LSP. This scenario in general presents problems for SUSY analyses in disentangling production processes.. Mixed-flavor production often preferentially gives same-sign leptons in the final state, compared to opposite-sign leptons from QCD pair production. Since the ATLAS atlas\_tdr and CMS cms\_tdr detectors at LHC will be able to identify lepton charge sign to better than $`10^4`$ accuracy, this characteristic of weak mixed-flavor production could provide a separation from the much greater rate of QCD production events. Our physics goal is to find and exploit characteristics of mixed-flavor squark production which would make these weak cross sections stand out against the much larger QCD squark production processes, providing a way to measure the weak vertices. ## II Stop-sbottom production at a hadron collider The best hope to look for mixed-flavor squark production is in the third generation, i.e. top and bottom squarks. The reasons for this are threefold. First, the largest first- and second-generation squark pair production mechanism at LHC is typically $`uu\stackrel{~}{u}_L\stackrel{~}{u}_L`$, which proceeds via a $`t`$-channel Majorana gluino and dominates over squark-antisquark production because of the very high initial-state valence $`u`$ quark luminosity. This state would decay predominately to same-sign leptons via $`\stackrel{~}{u}_L\chi _1^+d`$, with the pair of lightest charginos, $`\chi _1^+`$, decaying to same-sign leptons. Second, the first- and second-generation squarks are impossible to cleanly separate from each other and from gluino pair and gluino-squark production, due to the inability to tag light flavors. Tagging the $`b`$ jets in stop-sbottom mixed pairs would eliminate the first- and second-generation squark contributions, up to the level of fake $`b`$ tag rates. Third, because gluinos are Majorana, their decay chains give equal probability for either fermion sign SSlep-go . As a result, gluino pairs and squark-gluino production would be large sources of same-sign leptons. Tagging $`b`$ jets can likewise reduce this fake source of same-sign leptons, except that, depending on the spectrum, gluinos may decay into stops or sbottoms themselves. We will come back to these issues later as background considerations. For the third generation, the left- and right-handed squarks $`\stackrel{~}{q}_L`$, $`\stackrel{~}{q}_R`$ can mix, forming mass eigenstates $`\stackrel{~}{q}_1`$, $`\stackrel{~}{q}_2`$, with $`\stackrel{~}{q}_1`$ defined as lighter. The mixing is proportional to the Yukawa couplings of the SM fermions. The general mass matrix for stops is given by: $`M_{\stackrel{~}{t}}^2`$ $`=`$ $`\left(\begin{array}{cc}M_{\stackrel{~}{t}_L}^2& +m_t(A_t^{}\mu \mathrm{cot}\beta )\\ +m_t(A_t\mu ^{}\mathrm{cot}\beta )& M_{\stackrel{~}{t}_R}^2\end{array}\right)`$ (3) where $`\mathrm{tan}\beta `$ is the ratio of Higgs sector vacuum expectation values $`v_u/v_d`$, $`\mu `$ is the Higgsino mass parameter, $`A_t`$ is the top squark trilinear scalar coupling in the soft SUSY breaking potential, and $`M_{\stackrel{~}{t}_L}^2`$ $`=`$ $`m_{\stackrel{~}{t}_L}^2+M_Z^2\mathrm{cos}2\beta (I_{3t}s_W^2Q_t)+m_t^2,`$ (4) $`M_{\stackrel{~}{t}_R}^2`$ $`=`$ $`m_{\stackrel{~}{t}_R}^2+M_Z^2\mathrm{cos}2\beta s_W^2Q_t+m_t^2.`$ (5) where $`m_{\stackrel{~}{q}_L}`$, $`m_{\stackrel{~}{q}_R}`$ are the soft-breaking masses in the SUSY Lagrangian. For $`b`$ instead of $`t`$, replace $`\mathrm{cot}\beta `$ with $`\mathrm{tan}\beta `$. Because of this $`L`$-$`R`$ mixing in the third generation, the SUSY weak vertices $`W`$-$`\stackrel{~}{t}_i`$-$`\stackrel{~}{b}_j`$ are more complicated than the corresponding $`W`$-$`t`$-$`b`$ vertex in the SM: they contain not just the super-CKM angle $`\stackrel{~}{V}_{tb}`$, but also the mixing angles of the stops and sbottoms, which reflect the left-handed component of each squark. Expressed in terms of reduced parameters, the $`W`$-$`\stackrel{~}{t}_1`$-$`\stackrel{~}{b}_1`$ coupling is $`g_W\stackrel{~}{V}_{tb}\mathrm{cos}\theta _t\mathrm{cos}\theta _b`$, where the mixing angles are those that diagonalize Eq. 3 for stop and sbottom squarks, respectively. For $`\stackrel{~}{q}_2`$ instead of $`\stackrel{~}{q}_1`$, $`\mathrm{cos}\theta `$ is replaced by $`\mathrm{sin}\theta `$ to obtain the left-handed component. If SUSY were an exact symmetry, $`\stackrel{~}{V}_{tb}V_{tb}`$. This is a very good approximation even after SUSY breaking and evolution down to the electroweak scale Duncan:1983iq ; Donoghue:1983mx , although it can be altered if SUSY breaking is not flavor-blind. As in SM single-top production, there are three ways to produce mixed-flavor squarks at LHC: $`s`$-channel, $`t`$-channel, and $`W`$-associated production. The first two are shown in Figs. 1 and 2, respectively. The third is typically about 1/3 of the $`t`$-channel rate, but would be likely be experimentally indistinct from QCD sbottom pair production. We will therefore ignore this channel. The analogy to SM single-top production is not complete, since in the SM case the light $`b`$ quark mass allows it to be treated as an initial-state parton, which cannot happen here. The SM process also involves only one heavy state, whereas here we have two, leading to different kinematics. Finally, in the SUSY case, because the squarks are scalars, there is a four-point vertex which does not exist in the SM. Although the graphs of Fig. 2 are part of the real-emission QCD corrections to the basic process of Fig. 1, they behave very differently. As we will see, the $`t`$-channel cross sections are always much larger than the $`s`$-channel. This is a combination of the larger suppression from an $`s`$-channel propagator at large squark pair invariant mass, and initial-state gluon versus sea quark luminosity. To calculate cross sections, we use matrix elements generated by a new MSSM version of madgraph Stelzer:1994ta , called smadgraph SMG . Our calculations utilize CTEQ6L1 structure functions Pumplin:2002vw for the incoming protons at LHC, $`\sqrt{s}=14`$ TeV, and we choose the average of the two final-state squark masses as the factorization and renormalization scales. We begin by analyzing the cross sections for the MSSM parameter space benchmark point <sup>3</sup><sup>3</sup>3We use the SPS benchmarks Allanach:2002nj , designed to represent a number of canonical scenarios to aid exploratory phenomenology, only as examples for convenience, not as any suggestion that these scenarios are more likely to be realized in nature than any other. They should be regarded only as starting points for phenomenological investigation. SPS1a Allanach:2002nj . For this point, the squark masses are $`m_{\stackrel{~}{t}_1,\stackrel{~}{t}_2}=396,587`$ GeV and $`m_{\stackrel{~}{b}_1,\stackrel{~}{b}_2}=517,547`$ GeV. As seen in Table 1, the LO total cross sections for SPS1a are typically quite small, on the order of a few fb, although even this would produce many hundreds of events for the planned luminosity of 300 fb<sup>-1</sup> per experiment at the LHC, or thousands of events for the 3000 fb<sup>-1</sup> per experiment for the planned luminosity upgrade to the LHC, the SLHC Gianotti:2002xx (this would be, after all, a long-term measurement to determine the SUSY Lagrangian parameters, not a discovery channel). Note the asymmetry in $`\stackrel{~}{t}\stackrel{~}{b}^{}`$ v. $`\stackrel{~}{t}^{}\stackrel{~}{b}`$ production. It is due to the PDF asymmetry of the incoming protons, LHC being a $`p`$-$`p`$ collider, and the dominance (by approximately a factor of two) of initial-state up quarks over down quarks. Table 1 also reveals the left-handed and right-handed components of the stops and sbottoms. Since $`\stackrel{~}{b}_1`$ and $`\stackrel{~}{b}_2`$ are nearly degenerate, the relative ratio of their cross sections shows that it is $`\stackrel{~}{b}_1`$ which is mostly $`\stackrel{~}{b}_L`$. In contrast, while $`m_{\stackrel{~}{t}_2}m_{\stackrel{~}{t}_1}`$, the $`\stackrel{~}{t}_1`$ and $`\stackrel{~}{t}_2`$ cross sections are nearly the same, which shows that $`\stackrel{~}{t}_2`$ is more left-handed, although not by as much of a margin as in the sbottoms. In SUSY models where the mass parameters are unified at the GUT scale, it is a general feature that the lighter stop is more right-handed and the lighter sbottom is more left-handed. This comes about from the renormalization-group running of the mass parameters from the GUT to the TeV scale due to the different quantum numbers of top and bottom Drees:1995hj . We give the total $`\stackrel{~}{t}\stackrel{~}{b}`$ cross sections in Table 2, combining $`s`$\- and $`t`$-channel results and all $`\stackrel{~}{q}_1`$ and $`\stackrel{~}{q}_2`$ combinations, for all 10 SPS points, along with the stop and sbottom masses at each point. For comparison, we also show the total $`\stackrel{~}{q}_1\stackrel{~}{q}_1^{}`$+$`\stackrel{~}{q}_2\stackrel{~}{q}_2^{}`$ QCD production rates, calculated here with smadgraph, although this was first performed in Ref. Dawson:1983fw . The general feature is that rates are lower for heavier squarks, as expected due to phase space suppression, with an additional but non-obvious suppression from mixings. SPS1a has larger cross sections than most other SPS points, except for the light-stop scenario of SPS5. The questions are now, given some rather small rates, might any of these produce enough events to potentially be observable, what are the backgrounds in each potentially viable scenario, and what would measuring these cross sections tell us about SUSY? Before addressing the signal and background rates, we make a few comments on the significance of a potential stop-sbottom rate measurement. Ideally, one could measure the rate for each stop-sbottom pair independently and with good accuracy. One could then extract each of $`g_W\stackrel{~}{V}_{tb}\mathrm{cos}\theta _t\mathrm{cos}\theta _b`$, $`g_W\stackrel{~}{V}_{tb}\mathrm{cos}\theta _t\mathrm{sin}\theta _b`$, $`g_W\stackrel{~}{V}_{tb}\mathrm{sin}\theta _t\mathrm{cos}\theta _b`$ and $`g_W\stackrel{~}{V}_{tb}\mathrm{sin}\theta _t\mathrm{sin}\theta _b`$ independently. The sum of the squares of these is simply $`g_W^2\stackrel{~}{V}_{tb}^2`$, so this combination would in principle be a test of SUSY, as it would measure whether or not the weak $`W`$-$`\stackrel{~}{t}`$-$`\stackrel{~}{b}`$ vertex is of the same strength as the SM weak vertex $`W`$-$`t`$-$`b`$. Measurement of the weak coupling strength in only a single channel, or couple of channels, would be muddled by the mixing between left and right states. If this were all that was available, one would then argue that the measurement aids determination of the mixing angles. However, both of these ideas rely on the GUT-inspired model assumption of flavor-diagonal SUSY soft-breaking masses. This is not necessarily strongly motivated, as presumably flavor physics enters at the GUT scale if not before, and could well cause a deviation from such unifying assumptions. If the terms are not diagonal, then the mass matrices which diagonalize all the squarks mix the first two generations with the third, resulting in $`\stackrel{~}{V}_{tb}V_{tb}`$, and the stop-sbottom cross section(s) would help reveal the nature of the SUSY soft-breaking terms. Of course, in this case, all squarks would have some decay branching fraction to heavy flavor, which may be observable. This is obviously a much more complicated scenario, which we ignore in this first analysis for simplicity. ## III Projections for some SUSY scenarios Here we estimate the observability of stop-sbottom production in various scenarios at LHC or its luminosity upgrade, the SLHC. We begin with the largest cross section of all the SPS points, SPS5, then successively discuss SPS1a and various non-SPS points we formulate, ranging from small variations on the SPS points to arbitrary non-universal inputs scenarios motivated by phenomenology of the low-scale MSSM spectrum Drees:1995hj . ### III.1 SPS5 #### III.1.1 Signal We examine the SPS5 scenario first, as it has the largest stop-sbottom cross section of the SPS points. This is due primarily to the lighter stop being only 250 GeV, so the final state is not as phase space restricted as in most of the SPS scenarios, which typically have much more massive stops. In SPS5, we see that $`\stackrel{~}{t}_1`$ always decays as $`\stackrel{~}{t}_1b\chi _1^+`$, while the heavier stop decays this way $`16\%`$ of the time, $`57\%`$ to $`Z\stackrel{~}{t}_1Zb\chi _1^+`$, and $`20\%`$ to $`h\stackrel{~}{t}_1hb\chi _1^+`$, while the remaining fraction goes to $`t\chi _{1,2}^0`$. The lighter sbottom decays mostly to $`W^{}\stackrel{~}{t}_1`$ ($`77\%`$), with $`13\%`$ to $`t\chi _1^{}`$ and the remaining fraction to $`b\chi _2^0`$. The heavier sbottom has only a rare $`3\%`$ decay to $`t\chi _1^{}`$, preferring to go to $`b\chi _1^0`$ and $`W^{}\stackrel{~}{t}_1`$ approximately equally. We summarize the relevant sparticle masses, total widths at NLO in QCD, and branching fractions relevant for this analysis in Table 3. The decay combination we envision as being distinctive and indicative of stop-antisbottom production is $`\stackrel{~}{t}b\chi _1^+`$ and $`\stackrel{~}{b}^{}\overline{t}\chi _1^+`$, yielding $`b\overline{t}\chi _1^+\chi _1^+`$ (and its charge-conjugate for antistop-sbottom production). Having a top quark to reconstruct (hadronically) in the final state is, we feel, more distinctive than the $`b\overline{b}\chi _1^+\chi _1^{}W^\pm `$ state from sbottom decay to stop plus a $`W`$ boson, although this option could be explored as it would enlarge the signal sample considerably if accessible. We ignore decays to neutralinos, because they are charge-neutral and therefore cannot help distinguish $`\stackrel{~}{q}`$ from $`\stackrel{~}{q}^{}`$. For SPS5, and typically in most minimal-supergravity (mSUGRA) scenarios, the lightest chargino decays preferentially to a neutrino plus the lightest slepton, typically the lightest stau, $`\stackrel{~}{\tau }_1^\pm `$, as here. The stau in turn decays to the lightest SUSY particle (LSP), typically the lightest neutralino, plus a tau. The detector signature would then be $`b\overline{b}jj\tau ^\pm \tau ^\pm +/E_T`$. Both $`b`$ jets would be tagged by the vertex detector, with an efficiency of about $`50\%`$ each atlas\_tdr ; cms\_tdr . The taus can then in turn be identified by their leptonic decay ($`34\%`$ BR with $`ϵ_{\mathrm{ID}}=0.95`$) or their 1-prong hadronic decay ($`50\%`$ BR and $`ϵ_{\mathrm{ID}}=32\%`$Rainwater:1998kj . For both cases, the charge can be identified with uncertainty smaller than $`10^4`$, which is near-perfect, and sufficiently small to not worry about fake same-sign events from opposite-sign backgrounds. For the 1/4 of the time in the SPS5 scenario that each chargino decays instead to LSP plus $`W`$ boson, the leptonic decay of the latter would simply add to the final state, although it would alter the expected ratio of $`e`$ or $`\mu `$ to hadronic $`\tau `$ which would come from $`\stackrel{~}{\tau }\tau \chi _1^0`$ decays only. For LHC only, considering 600 fb<sup>-1</sup> by combining the results of two experiments, and combining stau and tau decay modes, we could expect about 550 signal events in the (++) channel before kinematic cuts, or 5500 at SLHC. This is promising, but must be put in the context of visibility above various backgrounds. #### III.1.2 Backgrounds There are multiple sources of backgrounds giving the same or similar final states. From the SM there is $`t\overline{t}W^+(t\overline{t}W^{})`$ production Maltoni:2002jr , which has a total cross section at leading order (LO) of 290(136) fb at LHC, or 2.47(1.16) fb after decay BRs to $`b\overline{b}jj\tau ^+\tau ^++/E_T`$, but before imposing any kinematic cuts. For a comparison with the case of signal taus decaying hadronically these are the relevant cross sections, and slightly larger than the signal, but within a factor of two. If the taus in the signal are to be observed in the lepton-hadron or dual lepton modes, then the SM background would be a factor 4 larger, although the signal would be a factor 5 larger. After cuts the $`t\overline{t}W`$ contribution is likely to be much smaller than the signal, as $`b`$ jets and taus and the reconstructed top quark from stop-sbottom decays will be produced much harder, i.e. with significantly more transverse momentum in the detector. More significantly, the SUSY signal is likely to have significantly more missing transverse momentum. It is probable that one can impose cuts such that most of the signal but very little of the SM background survives. However, this will require future complete decay simulation with kinematic cuts. QCD sbottom pair production can also give a similar final state: $`\stackrel{~}{b}_1W^{}\stackrel{~}{t}_1W^{}b\chi _1^+`$ and $`\stackrel{~}{b}_1^{}\overline{t}\chi _1^+`$, yielding $`b\overline{t}\chi _1^+\chi _1^++W^{}`$. The cross section for sbottoms pairs to this state of decays is 20.7 fb at LO, compared to 3.0 fb for the signal. This reveals a weakness of our proposal, which is that if in a given scenario the sbottom decays to stop with significant BR, then it could potentially fake the signal. However, this final state contains an extra $`W`$ boson, which could be vetoed in any of its decay modes. Previous applications of such a veto Wveto typically achieved more than an order of magnitude reduction, but this applied to processes such as top quark pair production, where the $`W`$ is produced with little kinematic boost. Here, the $`W`$ is the decay product of a 560 GeV object decaying to the $`W`$ (80 GeV) plus a 250 GeV object: the $`W`$ is likely to be boosted significantly, increasing the veto probability, although an exact calculation with kinematic cuts on the final-state $`W`$ decay products would be required to determine the precise rejection efficiency. Note that such a veto would also remove most of the $`W`$-associated signal channel. Hence, for this analysis we are jusitifed in not considering it. As for QCD stop pair production, while it is large, it arises almost exclusively from $`\stackrel{~}{t}_1\stackrel{~}{t}_1^{}`$ pairs, which decay $`100\%`$ to opposite-sign charginos. We find that $`\stackrel{~}{t}_2\stackrel{~}{t}_2^{}`$ decays to the signal signature are much less than a $`1\%`$ background to the signal rate. Finally, SUSY QCD production of a first- or second-generation squark plus a gluino, as well as gluino pairs, constitutes a potentially very large background. The reason is that squarks decay with large BR to quark plus chargino, and gluinos in SPS5 are sufficiently massive to decay to either top-stop or bottom-sbottom. Because gluinos are Majorana particles, they decay equally to particle-sparticle pairs of either sign permutation. This results in a very large cross section for same-sign lepton pairs from squark+gluino and gluino pair decays SSlep-go . To give an example, the largest squark-gluino cross section is $`\stackrel{~}{u}_L\stackrel{~}{g}`$, due to the initial state containing a valence $`u`$ quark <sup>4</sup><sup>4</sup>4Note that $`\stackrel{~}{q}_R`$ production is not a background as right-handed squarks cannot decay to charginos.. For SPS5 this alone has a cross section of 2300 fb at LO, with NLO QCD corrections being a factor 1.3 Beenakker:1996ch . The relevant decays are $`\stackrel{~}{u}_Ld\chi _1^+`$ ($`66\%`$) and $`\stackrel{~}{g}\overline{t}\stackrel{~}{t}_1,b\stackrel{~}{b}_1^{}`$ ($`27\%,10\%`$). These decays produce identifiable final-state content identical to the $`t`$-channel signal if the extra jet is observed: $`j_hb\overline{b}jj\tau ^+\tau ^++/E_T`$, where $`j_h`$ represents a very hard jet which arises from the $`d`$ quark. This jet is highly energetic due to the large mass difference between the squark and chargino, 680 GeV v. 230 GeV. Gluino pair production is 1600 fb at LO, with NLO QCD corrections of a factor 1.9 for the SPS5 gluino and squark masses Beenakker:1996ch . The background arises from one gluino decaying to quark plus squark (first- or second-generation only), while the other decays as in the squark-gluino case. The quark jet from the first gluino’s decay to quark plus squark, however, is typically soft for SPS5, so ultimately the signature is essentially the same as in squark-gluino production, but with a factor two for gluino decay combinatorics, and another factor two for decay to either up (down) or charm (strange) quark-squark pairs. We would not propose attempting to distinguish this background from the presence of the extra soft jet, as an additional soft jet is likely to arise also in a significant fraction of the signal simply due to QCD radiation. Thus, the heavy squark+gluino backgrounds are 3.8 and 1.9 pb for the (++) and ($``$) cases, respectively. These are a factor 240 larger than the signal after squark and gluino decays to chargino plus sbottom. Corrections to this from squark pair production with mistagged $`b`$ jets will be quite small, as the light quark rejection factor for $`b`$ jet fakes is 1/140. We do not consider this latter source further here. #### III.1.3 Differentiating signal and background Fortunately, this heavy squark+gluino background can be suppressed significantly by one of two means, either: (a) we take advantage of the kinematic characteristics inherent in $`\stackrel{~}{t}\stackrel{~}{b}^{}j`$, where the extra jet tends to be emitted at large rapidity and high transverse momentum in the detector, and require the hard jet from the squark decay to be similarly far forward; or (b) we veto the extra hard jet from the squark decay, since a sizeable fraction of the signal does not have a high-$`p_T`$ jet. We show the transverse momenta and rapidity distributions of the extra jet in signal $`\stackrel{~}{t}_1\stackrel{~}{b}_1^{}j`$ production in Fig. 3 for illustration. To be forward-tagged, we require $`p_T(j)>30`$ GeV and $`3<|\eta (j)|<5`$. This provides a suppression factor of 50, while 1/3 of the signal survives. This would make the background approximately a factor 15 larger than the signal - a mediocre ratio, but not immediately to be dismissed. For a jet veto, which is more speculative and would require far more in-depth investigation of additional QCD radiation in the signal, we veto if $`p_T(j)>40`$ GeV and $`|\eta (j)|<5`$. This veto brings the background down by a factor 113, while $`72\%`$ of the signal survives, for an overall signal-to-background (S/B) ratio of about 1/3.3, which is excellent. The S/B ratio is superior in the jet veto case because about twice as much signal survives, while the background rejection due to the hard squark decay jet is more than twice as good. Of course, our cuts values for these were not optimized, but serve as a useful rough guide for the moment. We also calculate the rates of $`\stackrel{~}{b}\stackrel{~}{b}^{}j`$ production using smadgraph, and find that there is a factor 7 rejection of this background if one requires a forward jet tag. This would likely be even better, as $`\stackrel{~}{b}\stackrel{~}{b}^{}jj`$ is also quite large. However, the rate for sbottom pair events to produce even one extra jet anywhere in the fiducial volume of the detectors and with $`p_T(j)>40`$ GeV is larger than either simple, parton-level $`\stackrel{~}{b}_i\stackrel{~}{b}_i^{}`$ rate at LO. While there are large uncertainties associated with such QCD calculations at LO, this is not entirely unreasonable, because the NLO corrections to sbottom pair production are quite large Beenakker:1997ut . What our result suggests is that the probability to produce an extra hard jet in squark pair events is extremely large, which would only enhance the rejection factor in the jet veto analysis. A similar result holds for squark+gluino production. This issue is sufficiently complicated to be beyond the scope of this paper, therefore is being investigated elsewhere XXnj . While we have not attempted to calculate the higher-order QCD (+1j) contributions to the signal, we anticipate they are much smaller than for the QCD processes, because no color is exchanged between the two incoming partons. Thus, additional radiation patterns should be moderate in comparison to QCD. This will be investigated in future work next . For now, our conclusion is that it is appropriate to ignore the sbottom pair backgrounds, as their rejection factors due to both hard QCD radiation and the extra $`W`$ boson certainly make them very small corrections to the extremely large squark+gluino background, as well as smaller than the signal. We will therefore focus on the squark+gluino backgrounds for our estimates of the feasibility of detecting the signal. #### III.1.4 Numerical estimates It is beyond the scope of this first-state work to calculate the signal and various backgrounds in full detail, with decays to a detector final state (even at the parton level) and applying kinematic cuts. We relegate this next logical state to future work next and here make only estimates of the upper bounds on signal cross sections. To make the estimates reasonable, we base them on branching ratios to the desired final state, as well as the known detector efficiencies for $`b`$ jet tagging and hadronic tau or lepton ID, depending on the exact final state. Our assumptions regarding the SLHC are that it collects 6000 fb<sup>-1</sup> (two experiments combined) and achieves the particle ID efficiencies discussed previously. We use our LO signal cross sections but NLO squark+gluino background numbers, since the corrections are large; this helps make our estimate conservative. We assume that the $`t\overline{t}W`$ background is eliminated completely, since it is the smallest to begin with, and ignore the sbottom pair background as discussed above. Thus, the numbers of signal and background events in the forward-tagged jet (jet veto) experiments are roughly 160(320) and 2350(1035). The statistical significance for the forward jet tag case is $`3\sigma `$, while for the jet veto option is $`10\sigma `$. This corresponds to an approximately $`12\%`$ uncertainty on the cross section from the veto analysis alone, half of that as the uncertainty on the weak $`W`$-$`\stackrel{~}{t}`$-$`\stackrel{~}{b}`$ coupling. Of course, this assumes that $`100\%`$ of the signal would be retained after kinematic cuts, which is obviously not correct. We would nevertheless expect a high retention rate because the final-state particles originate from the decays of very heavy objects, thus will appear centrally and at high transverse momentum in the detectors with a consequent high probability to pass even conservative kinematic acceptance cut requirements. We would anticipate loss of signal due to cuts to be less than a factor of two. Thus, our estimate is cause for optimism and invites more detailed investigation. One should realize, too, that the signal and background could behave differently with respect to the final-state kinematic cuts, so there is some additional systematic uncertainty at this stage as to the correct S/B ratio. ### III.2 SPS1a The benchmark point SPS1a has the next-largest stop-sbottom cross section, about half the rate of the SPS5, as shown in Table 2. This is primarily because the lightest stop, $`\stackrel{~}{t}_1`$, is more massive than in SPS5. The QCD squark and gluino pair background is much larger than in SPS5, in constrast, because the other squarks and gluino are less massive than in SPS5. However, production cross section numbers can be misleading because the different spectrum results in different BRs of the various states. In addition, the stop mixing is somewhat different, so that the $`\stackrel{~}{t}_2`$ cross sections are comparable to $`\stackrel{~}{t}_1`$, as we saw in Table 1. Here, $`\stackrel{~}{t}_1b\chi _1^+`$ only 2/3 of the time, compared to always in SPS5, while $`\stackrel{~}{b}_1\overline{t}\chi _1^+`$ occurs at three times the BR of SPS5. In SPS1a, there are also non-trivial contributions to the same final state from the sizeable $`\stackrel{~}{t}_2`$ rate. In the backgrounds, while there is far larger squark-gluino rate than in SPS5, the gluino BRs to bottom-sbottom and top-stop are a factor of several smaller. As with SPS5, for SPS1a we summarize in Table 4 the sparticle masses, total widths at NLO in QCD, and major branching fractions relevant for this analysis. We proceed as we did for the SPS5 scenario previously, constructing two possible analyses, one demanding the presence of a far-foward tagging jet, the other vetoing any additional hard jets, which the heavy squark decays would give. The rejection factors against the squark-gluino backgrounds are 40 and 80, respectively, which are about 4/5 of the factors in SPS5. Signal retention rates for both possible analyses are similar to those in SPS5. At this point, we estimate after all BRs, detector ID efficiencies, and including the NLO corrections for the heavy squark and gluino backgrounds, that the forward jet tag analysis could gather up to about 90 events in 6000 fb<sup>-1</sup>, against a background of about 3200 events, while the jet veto analysis could retain up to about 220 signal events against a background of 1600. Kinematic cuts on the final state particles will reduce this somewhat, but again we argue retention is likely to be highly efficient as the identifiable particles come from the decays of very heavy objects and thus tend to be highly boosted. While the tagged analysis is quite poor, the veto analysis already could achieve up to $`5\sigma `$, with up to a statistical uncertainty on the signal cross section of about $`20\%`$. One can improve on this by carefully considering the kinematic characteristics of both signal and background. The squark+gluino rate can be broken down into two major parts: $`\stackrel{~}{g}b\stackrel{~}{b}^{}`$, and $`\stackrel{~}{g}\overline{t}\stackrel{~}{t}`$; they are roughly equal. The bottom quark jet in gluino decays, however, is extremely soft, due to the small mass splitting between the gluino and the sbottom. The $`b`$ jet arising from stop decays in the signal, however, is quite hard, due to the rather large mass splitting between stop and chargino. For this first-stage study we implement one level of decays with smadgraph and calculate the rates for the signal process $`\stackrel{~}{t}\stackrel{~}{b}^{}b\chi _1^+\overline{t}\chi _1^+`$ as well as the background processes $`\stackrel{~}{u}_L\stackrel{~}{g}d\chi _1^+b\stackrel{~}{b}_1^{},d\chi _1^+\overline{t}\stackrel{~}{t}_1`$. We use the narrow width approximation for phase space, as the sparticle widths are typically a couple percent of their mass (see Table 4), but implement full matrix elements to retain spin angular correlations in the backgrounds, using total widths for the various particles as calculated at LO (to match the LO couplings in smadgraph) by the program sdecay Muhlleitner:2003vg ; we do consider overall BRs at NLO, also given by sdecay, since the corrections for the gluino can be large Beenakker:1996de . We show the $`p_T`$ distribution of the $`b`$ jet not arising from the top quark decay in Fig. 4. There is a dramatic difference in the kinematic behavior, as expected <sup>5</sup><sup>5</sup>5We examined this also for SPS5, but these kinematic features were not as dramatically different there.. We display two possible cuts, one which reduces the background by a factor 10 while retaining $`71\%`$ of the signal, the other which totally obliterates the background (factor 270) at the expense of fully half the signal. Note that imposing such a cut would also reduce the $`\stackrel{~}{g}\overline{t}\stackrel{~}{t}`$ rate by the approximately the same amount as the signal, as the $`b`$ jet not from top quark decay there comes from a stop as in the signal, so will display a similar kinematic behavior. A nice feature of the signal distribution is its relative flatness to high $`p_T`$: if the tail of the background distribution is wider than expected, one can adjust the cut to compensate for this at very little relative cost to the signal. Analogously for the other background contribution, $`\stackrel{~}{g}\overline{t}\stackrel{~}{t}`$, the top quark daughter will tend to be boosted more in the signal, where it was produced in conjuction with a fairly light chargino, whereas in the background its decay partner is a relatively heavy stop. Although the differences are not as dramatic as with the $`b`$ jet in stop v. gluino decays, we see in Fig. 5 that it is possible to find a cut which reduces the background by about an order of magnitude while retaining $`50\%`$ of the signal. This is a somewhat higher price to pay than in the gluino decay to bottom-sbottom, but it does improve the overall statistical significance, as well as the more important signal-to-background ratio. Again, imposing this cut would reduce the $`\stackrel{~}{g}b\stackrel{~}{b}^{}`$ portion of the background similar to the signal. Taking both cuts together, we estimate up to about 32(78) signal events could be retained in the forward-tag (jet veto) analysis, with a background of about 210(105) events. (Note that this is rather pessimistic, since the two cuts are not truly orthogonal, thus the actual signal retention rate should be higher.) This improves the jet veto case to a very respectable S/B ratio of 1/1.3, up to about $`7.6\sigma `$ statistical significance, and potentially up to a $`20\%`$ statistical uncertainty on the signal cross section. This is again idealized, but we would expect a high signal retention rate after cuts, as most of the final-state particles will be highly boosted and register well in the detectors. Again, as with SPS5 there is some systematic uncertainty on S/B that arises from not knowing whether the signal and background behave similarly for the final-state kinematic cuts. Further study will clarify this. ### III.3 General MSSM scenarios While the SPS scenarios constitue a useful framework in which to study SUSY phenomenology at upcoming collider experiments, we remind ourselves that they do not by any means fully represent the plethora of possible MSSM parameterizations. They are only starting points: the SUSY breaking schemes mSUGRA, gauge-mediated (GMSB) and anomaly-mediated (AMSB) all make broad assumptions about unification of input mass parameters at some high scale. While the motivations for doing so are elegant, they are not by any stretch of the imagination certain. Based on our experiences studying SPS5 and SPS1a, we can draw a few simple conclusions to guide exploration of parameter space with a more open mind. First, and most obviously, LHC has the potential to observe stop-sbottom production when the cross sections are large enough to produce a statistically useful number of events; the backgrounds in the SPS cases we examined so far can be suppressed to approximately the level of the signal. For the production of heavy objects, LHC is limited mostly by phase space. Thus, scenarios with lighter stops and sbottoms are more desireable. Second, very large backgrounds arise from first- and second-generation squark+gluino and gluino pair production, but they can also be heavily suppressed using additional kinematic information in the events. Scenarios with very heavy squarks and gluinos relative to stops and sbottoms are in some sense variants of SPS5 and highly likely to be accessible by comparison. Alternatively, if the gluino is lighter and cannot decay to stops or sbottoms, then this background completely disappears. The other squarks cannot become a sizeable background on their own, as their decays would not yield heavy flavor quarks (including a top quark) in the final state. One does have to worry about gluinos lighter than the sbottom, however, as the strong decay $`\stackrel{~}{b}b\stackrel{~}{g}`$ then becomes possible, which tends to dominate in BR over the weak decays to bottom quark plus chargino. Thus, while there is no background, there is little signal rate left to the same-sign charged lepton pair we advocate. Were $`b`$ jet-charge tagging to be possible with decent efficiency, then such a scenario might be accessible. One could also consider the possibility of identifying mixed-flavor production in this scenario by observing a single lepton from the stop decay, two $`b`$ jet tags, and a characteristic gluino, thus separating the signal from stop pair or sbottom pair production; but this is rather speculative and would require a completely different analysis. In searching for other viable scenarios, we do not rigorously check for known constraints from present data, except for the Higgs and chargino limits from LEP Barate:2003sz ; LEP-MSSM-H ; Clerbaux:2003gq ; Degrassi:2002fi , and of course the requirement of a neutral, colorless LSP. Nor do we attempt a thorough, systematic exploration of parameter space at this early stage. Our goal is simply to emphasize how different MSSM realizations could change expectations for the stop-sbottom signal qualitatively. An early warning: scenarios with a small NLSP-LSP mass difference can give final-state taus or leptons which are soft a large fraction of the time, presenting a rate problem after detector acceptance cuts. The exception to this is when the mass difference is so small that the NLSP is long-lived, as we will see below. * Lighter stop and sbottom via increased mixing. Starting with SPS1a and simply increasing the value of $`A_0`$ to -500 GeV, thus maintaining the unification of input mass parameters, results in about a $`70\%`$ cross section increase for the slightly lighter stops and sbottoms. However, the gluino mass doesn’t change, while its BR to top-stop increases by a factor three. Signal significane is likely to decrease slightly. * Heavier gluino via increased $`M_3`$. Starting with the low-energy effective SPS1a point but instead allowing non-universal values for $`m_{1/2}`$, setting $`M_{1,2,3}=100,200,500`$ GeV and $`A_{t,b}=2`$ TeV, we obtain a very heavy gluino, $`m_{\stackrel{~}{g}}=1154`$ GeV, while the stop and sbottom masses similar to those of SPS1a, and their total production cross sections about 1/3 the SPS1a rate. Now the squark+gluino backgrounds are only a few hundred fb to start, almost two orders of magnitude smaller, but with BRs to stops and sbottoms a factor of a few larger than SPS1a. With this spectrum the kinematic tricks we identified for SPS1a will not work as well, as the gluino-sbottom mass splitting is now several hundred GeV. However, the overall lowering of background from production rate, less the increased BRs, is about the same as the suppressed rate of SPS1a via kinematic differences. Such a scenario looks promising only if the gluino BR to stop or sbottom and loss of kinematic cut effectiveness does not grow at a faster rate than the cross section falls off due to phase space. * A variant of the above with $`M_{1,2,3}=150,300,1000`$ GeV and universal $`m_0=500`$ GeV at the TeV scale predicts a 970 GeV gluino, a light stop of 322 GeV and lightest sbottom of 475 GeV. The signal rates are similar to SPS5, with a smaller background than in SPS5, from both smaller production rates and smaller gluino decay rates to stops and sbottoms. We find a wide parameter space around this scenario which produces a similar spectrum. * Gluino lighter than the sbottoms. Starting at SPS1a but allowing non-universal masses, we lower $`m_{1/2}`$ for $`SU(3)`$ only to 100 GeV, set $`A_0=0`$ and $`m_{\stackrel{~}{t}_R}=m_{\stackrel{~}{b}_R}=150`$ GeV at the GUT scale. The gluino mass is now 271 GeV, while the lighter sbottom mass is 274 GeV. The lightest stop is 193 GeV; the heavier stop and sbottom are both under 400 GeV. The total of stop-sbottom cross sections is 160 fb, with no background from squark+gluino pairs. However, the lighter sbottom does not decay at all to chargino, rather mostly to $`b`$+LSP. The $`\stackrel{~}{b}_2`$ actually dominates the production rate, and it has a $`6\%`$ BR to chargino. The rate to $`b\overline{t}\chi _1^+\chi _1^+`$ is larger than at SPS1a by a small factor. The chargino is lighter than the stau in this scenario, which would decay like a $`W`$ boson, with an effective overall ID efficiency via its decays to $`e,\mu `$ about that same as a stau pair. Such scenarios look extremely promising. Other variations on this theme, with $`m_{\stackrel{~}{g}}m_{\stackrel{~}{b}}`$, are easy to find. Our qualitative survey suggests that parameterizations which give larger signal cross sections are often good, but also sometimes less viable, for a variety of reasons. Usually the culprit is either a dramatically larger BR of the gluino to a stop or sbottom, or the stop and sbottom BRs themselves squeeze the distinctive signal we propose. Analogously, smaller cross sections can sometimes be deceptively good, often for similar reasons. Stau coannihilation scenario We examine a stau-coannihilation scenario Ellis:1998kh ; Ellis:1999mm , inspired by the dark matter relic density Bertone:2004pz . We choose TeV-scale input parameters of $`M_{1,2,3}=100,100,800`$ GeV, $`m_0=800`$ GeV for the first two generations, $`m_{\stackrel{~}{\tau }}=127`$ GeV, $`m_{Q_L}=m_{t_R}=400`$ GeV, $`m_{b_R}=300`$ GeV and $`A_{t,b}=700`$ GeV. The stau mass obtained is 100 MeV above the LSP at 97 GeV, with a 101 GeV chargino - this is the principal features of this scenario. As a consequence, the stau is relatively long-lived. We selected other parameters to guarantee reasonably light third-generation squarks: the lighter stop and sbottom are 243 and 288 GeV, respectively, with their heavier partners at 544 and 406 GeV. The other squarks and gluino are somewhat heavier at 810 GeV. With regard to the colored sparticle sector, this point is not all that different from SPS5. The largest signal cross section is $`\stackrel{~}{t}_1\stackrel{~}{b}_2^{}`$ at 67 fb, as shown in Table 5. The stop decays $`100\%`$ to our desired chargino, and the sbottom 1/4 of the time. The chargino in turn decays exclusively to the lighter stau, which is long-lived. Herein lies the munificence: the detection efficiency for each stau will be extremely close to $`100\%`$, with perfect charge identification. Already, then, any such scenario will achieve at least a factor 4 detection efficiency improvement over any model where the stau decays promptly to a tau. Moreover, there is no SM background at all. For a jet tag strategy, we calculate 2.63 fb for the signal final state $`b\overline{t}\stackrel{~}{\tau }_1^+\stackrel{~}{\tau }_1^+`$ where we include a factor 2/3 for the hadronic BR of the top quark. For a jet veto strategy, we calculate 4.13 fb. Squark+gluino backgrounds are similar to SPS5, but with $`50\%`$ BR of gluino to bottom-sbottom and top-stop of the (++) charge sign, larger than SPS5. As implied, the gluino does not decay to any first- or second-generation squark plus quark, thus removing gluino pair production as a background. Rejection factors for the jet tag and jet veto scenarios are extremely good: 60 and 200, respectively. By the time NLO corrections, jet tag/veto rejection factors and BRs are taken into account, we find about 2.09 fb for the jet tag case and 0.61 fb using instead a jet veto. These cross sections are smaller than the signal. However, the real background comes from QCD sbottom pair production. Specifically, $`\stackrel{~}{b}_2\stackrel{~}{b}_2^{}`$ with a NLO cross section of approximately 1.9 pb, as $`\stackrel{~}{b}_2\stackrel{~}{b}_2^{}`$ cannot fake the signal at this point. After BRs, this is about 170 fb, although this is comparable to the sbottom pair rate for SPS5. The difference here is that the squark+gluino background is smaller, rather than significantly larger. Again we assume that there is an order of magnitude rejection of sbottom pairs via the extra $`W`$ boson, and another factor of 5 from a jet veto on the extra radiation. There is a large uncertainty associated with this, but as we will see, it is largely irrelevant for observational success at this point. Because of the anticipated large detection efficiency for long-lived staus, about a factor 5 over staus which would promptly decay to taus, our rough estimate of observability here will be for the LHC already, rather than the SLHC. Using a jet veto strategy we would anticipate approximately 620 double-$`b`$-tagged signal events (including hadronic top decay), on top of a total background of about 400 events; S/B would be better than 1/1. This would provide for a remarkable $`>30\sigma `$ detection and statistical uncertainty on the signal cross section to this final state of about $`5\%`$. Even if the background rejection factors due to $`W`$ or jet veto turn out to be off by an order of magnitude, this would still allow for a $`10\sigma `$ observation and up to an $`11\%`$ statistical uncertainty on the cross section. Needless to say, at SLHC this channel would rapidly become dominated by systematic errors, such as the QCD uncertainty on the signal theoretical cross section or the measurement of the $`\stackrel{~}{b}_2\stackrel{~}{b}_2^{}`$ rate and its BRs. However, because the dominant background is likely this squark pair rather than squark+gluino production, there is an additional handle one could apply to improve the situation: the signal has an approximately 2:1 asymmetry for (++) v. ($``$) production, due to the predominance of initial-state valence $`u`$ over valence $`d`$ partons. QCD sbottom pair production is totally symmetric, in contrast. Thus, by measuring the charge asymmetry of the final state, $$A=\frac{\sigma _{++}\sigma _{}}{\sigma _{++}+\sigma _{}},$$ (6) one could gain additional strong leverage over residual systematic uncertainty of the QCD background. GMSB scenario A particularly interesting scenario is gauge-mediated SUSY breaking (GMSB), which has a gravitino LSP. Because the decay of the NLSP would be of gravitational strength, the NLSP is typically long-lived, and if charged would be extraordinarily noticeable in experiment as a heavy charged object passing through the muon chambers. Because of this, the efficiency to detect SUSY events would be very close to $`100\%`$ if the NLSP is charged, completely eliminating SM backgrounds and retaining nearly the whole rate of stop-sbottom pairs, modulo the efficiency for $`b`$-tagging, needed to identify that the SUSY signal comes from third-generation squarks. SPS7 is a GMSB scenario with stau NLSP, but because of the very large stop and sbottom masses, there is practically no rate for mixed-flavor pairs at LHC. We instead look for parameterizations of GMSB where the stops and sbottoms are light enough to be produced at the LHC, and also where the gluino is slightly lighter than the sbottoms, so that the BR($`\stackrel{~}{g}b\stackrel{~}{b}`$) does not dominate. SUSY backgrounds are then composed only of light squarks and gluinos which decay to same-sign staus, with light jets mistagged as $`b`$ jets. The large suppression of such fakes may make such a scenario feasible for much smaller signal cross sections than in SPS1a or SPS5. We easily find an example of such a parameterization, by choosing $`M_{mes}=100`$ TeV, $`M_{SUSY}=20`$ TeV, $`\mathrm{tan}\beta =5`$, $`\mu <0`$, and the presence of 3 lepton and 2 quark messenger states. The latter is a somewhat odd choice, but is not restricted in any way. For these choices, $`m_{\stackrel{~}{t}_1,\stackrel{~}{t}_2,\stackrel{~}{b}_1,\stackrel{~}{b}_2}=350,410,360,370`$ GeV. Only the $`\stackrel{~}{b}_2`$ cross sections are of decent size: that for $`\stackrel{~}{t}_1\stackrel{~}{b}_2^{}`$ is about 1 fb and that for $`\stackrel{~}{t}_2\stackrel{~}{b}_2^{}`$ is about 3.6 fb, both obtained by requiring the forward-tagged jet as previously discussed. This would produce $`𝒪(50)`$ events at SLHC, including detector efficiencies but no kinematic cuts, which should be highly efficient given that the $`b`$ quarks would come from the decays of very heavy states. AMSB scenario Another interesting SUSY scenario is that of anomaly-mediation (AMSB), which typically possesses the characteristic of very small mass splitting between the LSP and NLSP, which are usually the lightest neutralino and chargino, respectively. The NLSP can exhibit dramatic vertex displacements in the detector of many centimeters before decaying. Because of this, SM backgrounds would be non-existent and the efficiency for capturing AMSB events would be extremely high, close to $`100\%`$, as in some GMSB scenarios. Charginos would decay to $`e`$ and $`\mu `$ each about 1/7 of the time, giving an overall clean, charge-identifiable state about $`8\%`$ of the time in stop-sbottom events. We do not explore AMSB parameter space widely, but find in general that it is difficult to obtain light-enough squark masses for LHC stop-sbottom production to be large enough to be useful, while at the same time avoiding the LEP constraints on the chargino mass. ## IV Outlook Our conclusions from this initial study relevant for LHC/SLHC are threefold. First, it does appear that there is a path to separately measuring weak production of mixed-flavor third-generation squark pairs in many SUSY scenarios at the LHC/SLHC. If further, more detailed studies bear this out, it represents a significant increase in the physics capability of LHC. Second, the interpretation of weak cross section measurements would depend strongly on what additional information is extractable from other production processes. The signals proposed here could represent a way to help disentangle third-generation squark mixing, at worst, or a potential measurement of the weak vertex and the flavor matrix that diagonalizes squarks, relative to the CKM matrix of the Standard Model. This would depend mostly what variety of SUSY is realized in nature, if at all: the relative spectrum, the overall scale of SUSY masses, and other unforeseeables. While hints of generational squark mixing may show up in QCD production channels, measuring the weak production vertex would greatly aid in testing the flavor-diagonal hypothesis and measuring deviations from it. Third, any scenario which has a long-lived NLSP is a boon for this measurement, as it would improve upon the detection efficiency of the signal by (at a minimum) a factor 4-5. Thus, any given model with squark and gluino masses that would appear to have a viable measurement of stop-sbottom production at SLHC would probably become viable at LHC if it also had a long-lived NSLP. Finally, we observe that our ability to pursue this interesting physics is limited only by cross section. Because squarks are typically heavy, copious production at future colliders could potentially open a new window onto these deeper questions, if SUSY is found in nature. One path is a next-generation (or beyond) linear collider ILC ; Group:2004sz , which could study the decays of squark pairs with great precision to help disentangle the mixing angles Bartl:1997yi . Another possible path would be a higher-energy hadron collider such as the proposed VLHC Baur:2002ka , which would have the ability to produce mixed-flavor squark pairs with cross sections typically two order of magnitude larger than at LHC. ### Acknowledgements We thank Tilman Plehn and Matt Strassler for highly useful discussions and critiques, Howie Baer for suggestions about additional SUSY backgrounds, and to Sally Dawson, Peter Zerwas and Lynne Orr for reviews of the manuscript. This research was supported in part by the U.S. Department of Energy under grant No. DE-FG02-91ER40685. ### Note added in proof: After submission of this work another paper appeared on the arXiv which also calculates the $`s`$-channel stop-sbottom cross sections Bozzi:2005sy . Their calculations for this channel agree with ours.
warning/0506/cond-mat0506588.html
ar5iv
text
# Imaging mesoscopic spin Hall flow: Spatial distribution of local spin currents and spin densities in and out of multiterminal spin-orbit coupled semiconductor nanostructures ## I Bond Spin Currents in Multiterminal Nanostructures: Landauer-Keldysh Approach The conservation of charge implies the continuity equation in quantum mechanics for the charge density $`\rho =e|\mathrm{\Psi }(𝐫)|^2`$ associated with a given wave function $`\mathrm{\Psi }(𝐫)`$ $$\frac{\rho }{t}+𝐣=0,$$ (1) from which one can extract the expression for the charge current density $$𝐣=e\mathrm{Re}[\mathrm{\Psi }^{}(𝐫)\widehat{𝐯}\mathrm{\Psi }(𝐫)].$$ (2) This can be viewed as the quantum-mechanical expectation value \[in the state $`\mathrm{\Psi }(𝐫)`$\] of the charge current density operator $$\widehat{𝐣}=e\frac{\widehat{n}(𝐫)\widehat{𝐯}+\widehat{𝐯}\widehat{n}(𝐫)}{2},$$ (3) which follows from the classical charge current density $`𝐣=en(𝐫)𝐯`$ via quantization procedure where the particle density $`n(𝐫)`$ and the velocity $`𝐯`$ are replaced by the corresponding operators and symmetrized to ensure that $`\widehat{𝐣}`$ is a Hermitian operator. baranger1989a In SO coupled systems $`\widehat{𝐣}`$ acquires extra terms since the velocity operator $`i\mathrm{}\widehat{𝐯}=[\widehat{𝐫},\widehat{H}]`$ is modified by the presence of SO terms in the Hamiltonian $`\widehat{H}`$. For example, for the effective mass Rashba Hamiltonian of a finite-size 2DEG structure (in the $`xy`$-plane) $$\widehat{H}=\frac{\widehat{𝐩}^2}{2m^{}}+\frac{\alpha }{\mathrm{}}\left(\widehat{p}_y\widehat{\sigma }_x\widehat{p}_x\widehat{\sigma }_y\right)+V_{\mathrm{conf}}(x,y),$$ (4) the velocity operator is $$\widehat{𝐯}=\frac{\widehat{𝐩}}{m^{}}\frac{\alpha }{\mathrm{}}(\widehat{\sigma }_y𝐞_x\widehat{\sigma }_x𝐞_y),$$ (5) where $`𝐞_x`$ and $`𝐞_y`$ are the unit vectors along the $`x`$ and the $`y`$-axis, respectively. Here $`\widehat{𝐩}=(\widehat{p}_x,\widehat{p}_y)`$ is the momentum operator in 2D space, $`\widehat{𝝈}=(\widehat{\sigma }_x,\widehat{\sigma }_y,\widehat{\sigma }_z)`$ is the vector of the Pauli spin matrices, $`\alpha `$ is the strength of the Rashba SO coupling rashba\_review ; winkler\_book arising due to the structure inversion asymmetry, pfeffer1998a and $`V_{\mathrm{conf}}(x,y)`$ is the transverse confining potential. In contrast to the charge continuity equation Eq. (1), the analogous continuity equation for the spin density $`\rho _s^i=\frac{\mathrm{}}{2}[\mathrm{\Psi }^{}(𝐫)\widehat{\sigma }_i\mathrm{\Psi }(𝐫)]`$ $$\frac{\rho _s^i}{t}+𝓙^i=S_s^i,$$ (6) contains the spin current density $$𝓙^i=\frac{\mathrm{}}{2}\mathrm{\Psi }^{}(𝐫)\frac{\sigma _i\widehat{𝐯}+\widehat{𝐯}\sigma _i}{2}\mathrm{\Psi }(𝐫),$$ (7) as well as a non-zero spin source $$S_s^i=\frac{\mathrm{}}{2}\mathrm{Re}\left(\mathrm{\Psi }^{}(𝐫)\frac{i}{\mathrm{}}[\widehat{H},\widehat{\sigma }_i]\mathrm{\Psi }(𝐫)\right).$$ (8) The non-zero $`S_s^i0`$ term reflects non-conservation of spin in the presence of SO couplings. Thus, the plausible Hermitian operator of the spin current density rashba\_review $$\widehat{𝒥}_k^i=\frac{\mathrm{}}{2}\frac{\sigma _i\widehat{v}_k+\widehat{v}_k\sigma _i}{2}$$ (9) is a well-defined quantity (a tensor with nine components) only when the velocity operator is spin independent, as encountered in many metal spintronic devices. brataas2001a Such lack of physical justification for Eq. (9) in SO coupled systems leads to an arbitrariness murakami2003a in the definition of the spin current density employed in recent intrinsic spin Hall studies, niu2005a thereby casting a doubt on the experimental relevance of the quantitative predictions murakami2003a ; sinova2004a for the spin Hall conductivity $`\sigma _{sH}=𝒥_y^z/E_x`$ computed as the linear response to the applied longitudinal electric field $`E_x`$ penetrating an infinite SO coupled (perfect) semiconductor crystal. To obtain the spatial profiles of spin and charge current densities in finite-size samples of arbitrary shape attached to many probes, it is advantageous to represent the spin-dependent Hamiltonian and the corresponding charge and spin current density operators in the local orbital basis. todorov2002a ; caroli1971a ; nonoyama1998a ; cresti2003a For example, in such representation the Rashba Hamiltonian can be recast in the following form nikolic\_accumulation $`\widehat{H}={\displaystyle \underset{𝐦\sigma }{}}\epsilon _𝐦I_s\widehat{c}_{𝐦\sigma }^{}\widehat{c}_{𝐦\sigma }+{\displaystyle \underset{\mathrm{𝐦𝐦}^{}\sigma \sigma ^{}}{}}\widehat{c}_{𝐦\sigma }^{}t_{\mathrm{𝐦𝐦}^{}}^{\sigma \sigma ^{}}\widehat{c}_{𝐦^{}\sigma ^{}},`$ (10) where hard wall boundary conditions account for confinement on the lattice $`L_x\times L_y`$ with the lattice spacing $`a`$. Here $`\widehat{c}_{𝐦\sigma }^{}`$ ($`\widehat{c}_{𝐦\sigma }`$) is the creation (annihilation) operator of an electron at the site $`𝐦=(m_x,m_y)`$. While this Hamiltonian is of tight-binding type, its off-diagonal elements are non-trivial $`2\times 2`$ Hermitian matrices $`𝐭_{𝐦^{}𝐦}=(𝐭_{\mathrm{𝐦𝐦}^{}})^{}`$ in the spin space. The on-site potential $`\epsilon _𝐦`$ describes any static local potential, such as the electrostatic potential due to the applied voltage or the disorder simulated via a uniform random variable $`\epsilon _𝐦[W/2,W/2]`$. The generalized nearest neighbor hopping $`t_{\mathrm{𝐦𝐦}^{}}^{\sigma \sigma ^{}}=(𝐭_{\mathrm{𝐦𝐦}^{}})_{\sigma \sigma ^{}}`$ accounts for the Rashba coupling $`𝐭_{\mathrm{𝐦𝐦}^{}}=\{\begin{array}{cc}t_\mathrm{o}I_sit_{\mathrm{SO}}\widehat{\sigma }_y& (𝐦=𝐦^{}+𝐞_x)\\ t_\mathrm{o}I_s+it_{\mathrm{SO}}\widehat{\sigma }_x& (𝐦=𝐦^{}+𝐞_y)\end{array},`$ (13) through the SO hopping parameter $`t_{\mathrm{SO}}=\alpha /2a`$ ($`I_s`$ is the unit $`2\times 2`$ matrix in the spin space). The direct correspondence between the continuous Eq. (4) and the lattice Hamiltonian Eq. (10) is established by using $`t_\mathrm{o}=\mathrm{}^2/2m^{}a^2`$ for the orbital hopping and by selecting the Fermi energy ($`E_F=3.8t_\mathrm{o}`$ in the rest of the paper) at which zero-temperature quantum transport takes place close to the bottom of the band $`E_b=4.0t_\mathrm{o}`$ to ensure that injected quasiparticles have quadratic and isotropic energy-momentum dispersion which characterizes the Hamiltonians in effective mass approximation. Using the effective mass and conduction bandwidth of the semiconductor heterostructures measured in experiments, nitta1997a one can interpret the parameters of the lattice Hamiltonian as having the typical values $`a3`$ nm for the lattice spacing while the Rashba SO hopping is of the order of $`t_{\mathrm{SO}}/t_\mathrm{o}0.01`$. The usage of the second quantized notation in Eq. 10 facilitates the introduction of Keldysh Green function keldysh1965a expressions for the nonequilibrium expectation values. caroli1971a ; nonoyama1998a We imagine that at time $`t^{}=\mathrm{}`$ the sample and the leads are not connected, while the left and the right longitudinal lead of a four-probe device are in their own thermal equilibrium with the chemical potentials $`\mu _L`$ and $`\mu _R`$, respectively, where $`\mu _L=\mu _R+eV`$. The adiabatic switching of the hopping parameter connecting the leads and the sample generates time evolution of the density matrix of the structure. caroli1971a The physical quantities are obtained as the nonequilibrium statistical average $`\mathrm{}`$ (with respect to the density matrix keldysh1965a at time $`t^{}=0`$) of the corresponding quantum-mechanical operators expressed in terms of $`\widehat{c}_{𝐦\sigma }^{}`$ and $`\widehat{c}_{𝐦\sigma }`$. This will lead to the expressions of the type $`\widehat{c}_{𝐦\sigma }^{}\widehat{c}_{𝐦\sigma ^{}}`$, which define the lesser Green function caroli1971a ; nonoyama1998a $`\widehat{c}_{𝐦\sigma }^{}\widehat{c}_{𝐦^{}\sigma ^{}}`$ $`=`$ $`{\displaystyle \frac{\mathrm{}}{i}}G_{𝐦^{}𝐦,\sigma ^{}\sigma }^<(\tau =0)`$ (14) $`=`$ $`{\displaystyle \frac{1}{2\pi i}}{\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑EG_{𝐦^{}𝐦,\sigma ^{}\sigma }^<(E).`$ Here we utilize the fact that the two-time correlation function \[$`\widehat{c}_{𝐦\sigma }(t)=e^{i\widehat{H}t/\mathrm{}}\widehat{c}_{𝐦\sigma }e^{i\widehat{H}t/\mathrm{}}`$\] $$G_{\mathrm{𝐦𝐦}^{},\sigma \sigma ^{}}^<(t,t^{})\frac{i}{\mathrm{}}\widehat{c}_{𝐦^{}\sigma ^{}}^{}(t^{})\widehat{c}_{𝐦\sigma }(t),$$ (15) depends only on $`\tau =tt^{}`$ in stationary situations, so it can be Fourier transformed to energy $$G_{\mathrm{𝐦𝐦}^{},\sigma \sigma ^{}}^<(\tau )=\frac{1}{2\pi \mathrm{}}_{\mathrm{}}^{\mathrm{}}𝑑EG_{\mathrm{𝐦𝐦}^{},\sigma \sigma ^{}}^<(E)e^{iE\tau /\mathrm{}},$$ (16) which will be utilized for steady-state transport studied here. We use the notation where $`𝐆_{\mathrm{𝐦𝐦}^{}}^<`$ is a $`2\times 2`$ matrix in the spin space whose $`\sigma \sigma ^{}`$ element is $`G_{\mathrm{𝐦𝐦}^{},\sigma \sigma ^{}}^<`$. ### I.1 Bond charge currents in SO coupled systems #### I.1.1 Bond charge-current operator Using the charge conservation expressed through the familiar continuity equation Eq. (1) yields a uniquely determined bond charge current operator for quantum systems described on a lattice by a tight-binding-type of Hamiltonian Eq. (10). That is, the Heisenberg equation of motion $`{\displaystyle \frac{d\widehat{N}_𝐦}{dt}}={\displaystyle \frac{1}{i\mathrm{}}}[\widehat{N}_𝐦,\widehat{H}].`$ (17) for the electron number operator $`\widehat{N}_𝐦`$ on site $`𝐦`$, $`\widehat{N}_𝐦{\displaystyle \underset{\sigma =,}{}}\widehat{c}_{𝐦\sigma }^{}\widehat{c}_{𝐦\sigma },`$ (18) leads to the charge continuity equation on the lattice $`e{\displaystyle \frac{d\widehat{N}_𝐦}{dt}}+{\displaystyle \underset{k=x,y}{}}\left(\widehat{J}_{𝐦,𝐦+𝐞_k}\widehat{J}_{𝐦𝐞_k,𝐦}\right)=0.`$ (19) This equation introduces the bond charge-current operator baranger1989a ; todorov2002a $`\widehat{J}_{\mathrm{𝐦𝐦}^{}}`$ which describes the particle current from site $`𝐦`$ to its nearest neighbor site $`𝐦^{}`$ (the ‘bond’ terminology is supported by a picture where current between two sites is represented by a bundle of flow lines bunched together along a line joining the two sites). Thus, the spin-dependent Hamiltonian Eq. (10) containing the non-trivial $`2\times 2`$ matrix hopping defines the bond charge current operator $`\widehat{J}_{\mathrm{𝐦𝐦}^{}}=_{\sigma \sigma ^{}}\widehat{J}_{\mathrm{𝐦𝐦}^{}}^{\sigma \sigma ^{}}`$ which can be viewed as the sum of four different spin-resolved bond charge-current operators $$\widehat{J}_{\mathrm{𝐦𝐦}^{}}^{\sigma \sigma ^{}}=\frac{e}{i\mathrm{}}\left[\widehat{c}_{𝐦^{}\sigma ^{}}^{}t_{𝐦^{}𝐦}^{\sigma ^{}\sigma }\widehat{c}_{𝐦\sigma }\text{h.c.}\right],$$ (20) where h.c. stands for the Hermitian conjugate of the first term. In particular, for the case of $`t_{\mathrm{𝐦𝐦}^{}}^{\sigma \sigma ^{}}`$ being determined by the Rashba SO interaction Eq. (13), we can decompose the bond charge current operator into two terms $$\widehat{J}_{\mathrm{𝐦𝐦}^{}}=\widehat{J}_{\mathrm{𝐦𝐦}^{}}^{\mathrm{kin}}+\widehat{J}_{\mathrm{𝐦𝐦}^{}}^{\mathrm{SO}},$$ (21) which have transparent physical interpretation. The first term $$\widehat{J}_{\mathrm{𝐦𝐦}^{}}^{\mathrm{kin}}=\underset{\sigma }{}\frac{et_\mathrm{o}}{i\mathrm{}}\left[\widehat{c}_{𝐦^{}\sigma }^{}\widehat{c}_{𝐦\sigma }\text{h.c.}\right]$$ (22) can be denoted as “kinetic” since it originates only from the kinetic energy $`t_\mathrm{o}`$ and does not depend on the SO coupling energy $`t_{\mathrm{SO}}`$, while the second term $`\widehat{J}_{\mathrm{𝐦𝐦}^{}}^{\mathrm{SO}}`$ $`=`$ $`\{\begin{array}{cc}{\displaystyle \frac{4et_{\mathrm{SO}}}{\mathrm{}^2}}\widehat{S}_{\mathrm{𝐦𝐦}^{}}^y& (𝐦^{}=𝐦+𝐞_x)\\ +{\displaystyle \frac{4et_{\mathrm{SO}}}{\mathrm{}^2}}\widehat{S}_{\mathrm{𝐦𝐦}^{}}^x& (𝐦^{}=𝐦+𝐞_y)\end{array}`$ (25) $`=`$ $`{\displaystyle \frac{4et_{\mathrm{SO}}}{\mathrm{}^2}}\left((𝐦^{}𝐦)\times \widehat{𝐒}_{\mathrm{𝐦𝐦}^{}}\right)_z`$ (26) is the additional contribution to the intersite current flow due to non-zero Rashba SO hopping $`t_{\mathrm{SO}}`$. Here we also introduce the “bond spin-density” operator $`\widehat{𝐒}_{\mathrm{𝐦𝐦}^{}}`$ $`=`$ $`{\displaystyle \frac{\mathrm{}}{4}}{\displaystyle \underset{\alpha \beta }{}}\left[\widehat{c}_{𝐦^{}\alpha }^{}\widehat{𝝈}_{\alpha \beta }\widehat{c}_{𝐦\beta }+\text{h.c.}\right],`$ (27) defined for the “bond” connecting the sites $`𝐦`$ and $`𝐦^{}`$, which reduces to the usual definition of the local spin density operator for $`𝐦^{}=𝐦`$ \[see Eq. (39)\]. #### I.1.2 Nonequilibrium bond charge current The formalism of bond charge currents yields the physically measurable topinka2003a local charge current within the sample as the quantum-statistical average $`\mathrm{}`$ (with respect to a density matrix that has evolved over sufficiently long time so that nonequilibrium state and all interactions are fully established) of the bond charge-current operator in the nonequilibrium state, caroli1971a ; nonoyama1998a ; cresti2003a $`\widehat{J}_{\mathrm{𝐦𝐦}^{}}`$ $`=`$ $`{\displaystyle \underset{\sigma \sigma ^{}}{}}\widehat{J}_{\mathrm{𝐦𝐦}^{}}^{\sigma \sigma ^{}},`$ (28) $`\widehat{J}_{\mathrm{𝐦𝐦}^{}}^{\sigma \sigma ^{}}`$ $`=`$ $`{\displaystyle \frac{e}{\mathrm{}}}{\displaystyle _{\mathrm{}}^{\mathrm{}}}{\displaystyle \frac{dE}{2\pi }}[t_{𝐦^{}𝐦}^{\sigma ^{}\sigma }G_{\mathrm{𝐦𝐦}^{},\sigma \sigma ^{}}^<(E)`$ (29) $`t_{\mathrm{𝐦𝐦}^{}}^{\sigma \sigma ^{}}G_{𝐦^{}𝐦,\sigma ^{}\sigma }^<(E)],`$ where we utilize Eq. (14) to express the local charge current in terms of the nonequilibrium lesser Green function. The spin-resolved bond charge current in (29) describes flow of charges which start as spin $`\sigma `$ electrons at the site $`𝐦`$ and end up as a spin $`\sigma ^{}`$ electrons at the site $`𝐦^{}`$ where possible spin-flips $`\sigma \sigma ^{}`$ (instantaneous or due to precession) are caused by spin-dependent interactions. The decomposition of the bond charge-current operator into kinetic and SO terms in Eq. (21) leads to a Green function expression for the corresponding nonequilibrium bond charge currents $`\widehat{J}_{\mathrm{𝐦𝐦}^{}}=\widehat{J}_{\mathrm{𝐦𝐦}^{}}^{\mathrm{kin}}+\widehat{J}_{\mathrm{𝐦𝐦}^{}}^{\mathrm{SO}}`$ with kinetic and SO terms given by $`\widehat{J}_{\mathrm{𝐦𝐦}^{}}^{\mathrm{kin}}`$ $`=`$ $`{\displaystyle \frac{et_\mathrm{o}}{\mathrm{}}}{\displaystyle _{\mathrm{}}^{\mathrm{}}}{\displaystyle \frac{dE}{2\pi }}\mathrm{Tr}_s\left[𝐆_{\mathrm{𝐦𝐦}^{}}^<(E)𝐆_{𝐦^{}𝐦}^<(E)\right],`$ $`\widehat{J}_{\mathrm{𝐦𝐦}^{}}^{\mathrm{SO}}`$ $`=`$ $`{\displaystyle \frac{et_{\mathrm{SO}}}{\mathrm{}}}{\displaystyle _{\mathrm{}}^{\mathrm{}}}{\displaystyle \frac{dE}{2\pi i}}`$ $`\times \mathrm{Tr}_s\left[\left((𝐦^{}𝐦)\times 𝝈\right)_z\left(𝐆_{\mathrm{𝐦𝐦}^{}}^<(E)+𝐆_{𝐦^{}𝐦}^<(E)\right)\right].`$ Note, however, that kinetic term is also influenced by the SO coupling through $`𝐆^<`$. In the absence of the SO coupling, Eq. (I.1.2) vanishes and the bond charge current reduces to the standard expression. caroli1971a ; nonoyama1998a ; cresti2003a The trace $`\mathrm{Tr}_s`$ here is performed in the spin Hilbert space. Similarly, we can also obtain the nonequilibrium local charge density in terms of $`𝐆^<`$ $`e\widehat{N}_𝐦`$ $`=`$ $`e{\displaystyle \underset{\alpha =,}{}}\widehat{c}_{𝐦\alpha }^{}\widehat{c}_{𝐦\alpha }`$ (32) $`=`$ $`{\displaystyle \frac{e}{2\pi i}}{\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑E{\displaystyle \underset{\sigma }{}}G_{\mathrm{𝐦𝐦},\sigma \sigma }^<(E)`$ $`=`$ $`{\displaystyle \frac{e}{2\pi i}}{\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑E\mathrm{Tr}_\mathrm{s}[𝐆_{\mathrm{𝐦𝐦}}^<(E)],`$ which is the statistical average value of the corresponding operator Eq. (18). ### I.2 Bond Spin Currents in SO Coupled Systems #### I.2.1 Bond spin-current operator To mimic the plausible definition of the spin-current density operator $`𝒥_k^i`$ in Eq. (9), we can introduce the bond spin-current operator for the spin-$`S_i`$ component as the symmetrized product of the spin-$`\frac{1}{2}`$ operator $`\mathrm{}\widehat{\sigma }_i/2`$ ($`i=x,y,z`$) and the bond charge-current operator from Eq. (19) $`\widehat{J}_{\mathrm{𝐦𝐦}^{}}^{S_i}`$ $``$ $`{\displaystyle \underset{\alpha \beta }{}}{\displaystyle \frac{1}{4i}}\left[\widehat{c}_{𝐦^{}\beta }^{}\{\widehat{\sigma }_i,𝐭_{𝐦^{}𝐦}\}_{\beta \alpha }\widehat{c}_{𝐦\beta }\text{h.c.}\right].`$ (33) By inserting the hopping matrix $`𝐭_{𝐦^{}𝐦}`$ Eq. (13) of the lattice SO Hamiltonian into this expression we obtain its explicit expression for the Rashba SO coupled system $`\widehat{J}_{\mathrm{𝐦𝐦}^{}}^{S_i}`$ $`=`$ $`{\displaystyle \frac{t_\mathrm{o}}{2i}}{\displaystyle \underset{\alpha \beta }{}}\left(\widehat{c}_{𝐦^{}\beta }^{}\left(\sigma _i\right)_{\beta \alpha }\widehat{c}_{𝐦\alpha }\text{h.c.}\right)`$ (34) $`+{\displaystyle \frac{t_{\mathrm{SO}}}{2}}\widehat{N}_{\mathrm{𝐦𝐦}^{}}\left(𝐞_i\times (𝐦^{}𝐦)\right)_z,`$ which can be considered as the lattice version of Eq. (9). Here $`\widehat{N}_{\mathrm{𝐦𝐦}^{}}`$ is the bond electron-number operator $`\widehat{N}_{\mathrm{𝐦𝐦}^{}}{\displaystyle \frac{1}{2}}{\displaystyle \underset{\sigma }{}}(\widehat{c}_{𝐦^{}\sigma }^{}\widehat{c}_{𝐦\sigma }+\mathrm{h}.\mathrm{c}.),`$ (35) which reduces to the standard electron-number operator Eq. (18) for $`𝐦^{}=𝐦`$. #### I.2.2 Nonequilibrium bond spin current Similarly to the case of the nonequilibrium bond charge current in Sec. I.1.2, the nonequilibrium statistical average of the bond spin-current operator Eq. (34) can be expressed using the lesser Green function $`𝐆^<`$ as $`\widehat{J}_{\mathrm{𝐦𝐦}^{}}^{S_i}`$ $`=`$ $`\widehat{J}_{\mathrm{𝐦𝐦}^{}}^{S_i(\mathrm{kin})}+\widehat{J}_{\mathrm{𝐦𝐦}^{}}^{S_i(\mathrm{so})}`$ (36) $`\widehat{J}_{\mathrm{𝐦𝐦}^{}}^{S_i(\mathrm{kin})}`$ $`=`$ $`{\displaystyle \frac{t_\mathrm{o}}{2}}{\displaystyle _{\mathrm{}}^{\mathrm{}}}{\displaystyle \frac{dE}{2\pi }}`$ (37) $`\times \mathrm{Tr}\left[\sigma _i\left(𝐆_{\mathrm{𝐦𝐦}^{}}^<(E)𝐆_{𝐦^{}𝐦}^<(E)\right)\right],`$ $`\widehat{J}_{\mathrm{𝐦𝐦}^{}}^{S_i(\mathrm{SO})}`$ $`=`$ $`\left(𝐞_i\times (𝐦^{}𝐦)\right)_z{\displaystyle \frac{t_{\mathrm{SO}}}{2}}{\displaystyle _{\mathrm{}}^{\mathrm{}}}{\displaystyle \frac{dE}{2\pi i}}`$ (38) $`\times \mathrm{Tr}\left[𝐆_{\mathrm{𝐦𝐦}^{}}^<(E)+𝐆_{𝐦^{}𝐦}^<(E)\right].`$ Here we also find two terms which can be interpreted as the kinetic and the SO contribution to the bond spin current crossing from site $`𝐦`$ to site $`𝐦^{}`$. However, we emphasize that such SO contribution to the spin-$`S_z`$ bond current is identically equal to zero, which simplifies the expression for this component to Eq. (37) studied in the rest of the paper as the primary spin current response in the spin Hall effect. #### I.2.3 Local Spin density and its continuity equation The local spin density in the lattice models is determined by the local spin operator $`\widehat{𝐒}_𝐦=(\widehat{S}_𝐦^x,\widehat{S}_𝐦^y,\widehat{S}_𝐦^z)`$ at site $`𝐦`$ defined by $`\widehat{𝐒}_𝐦={\displaystyle \frac{\mathrm{}}{2}}{\displaystyle \underset{\alpha \beta }{}}\widehat{c}_{𝐦\alpha }^{}𝝈_{\alpha \beta }\widehat{c}_{𝐦\beta }.`$ (39) The Heisenberg equation of motion for each component $`\widehat{𝐒}_i`$ ($`i=x,y,z`$) of the spin density operator $$\frac{d\widehat{S}_𝐦^i}{dt}=\frac{1}{i\mathrm{}}[\widehat{S}_𝐦^i,\widehat{H}]$$ (40) can be written in the following form $`{\displaystyle \frac{d\widehat{S}_𝐦^i}{dt}}+{\displaystyle \underset{k=x,y}{}}\left(\widehat{J}_{𝐦,𝐦+𝐞_k}^{S_i}\widehat{J}_{𝐦𝐞_k,𝐦}^{S_i}\right)=\widehat{F}_𝐦^{S_i},`$ (41) where $`\widehat{J}_{\mathrm{𝐦𝐦}^{}}^{S_i}`$ is the bond spin-current operator given by Eq. (33) so that the second term on the lefthand side of Eq. (41) corresponds to the “divergence” of the bond spin current on site $`𝐦`$. Here, in analogy with Eq. (8), we also find the lattice version of the spin source operator $`\widehat{F}_𝐦^{S_i}`$ whose explicit expression is given by $`\widehat{F}_𝐦^{S_x}`$ $`=`$ $`{\displaystyle \frac{t_{\mathrm{SO}}}{t_\mathrm{o}}}\left(\widehat{J}_{𝐦,𝐦+𝐞_x}^{S_z}+\widehat{J}_{𝐦𝐞_x,𝐦}^{S_z}\right),`$ (42) $`\widehat{F}_𝐦^{S_y}`$ $`=`$ $`{\displaystyle \frac{t_{\mathrm{SO}}}{t_\mathrm{o}}}\left(\widehat{J}_{𝐦,𝐦+𝐞_y}^{S_z}+\widehat{J}_{𝐦𝐞_y,𝐦}^{S_z}\right),`$ (43) $`\widehat{F}_𝐦^{S_z}`$ $`=`$ $`{\displaystyle \frac{t_{\mathrm{SO}}}{t_\mathrm{o}}}(\widehat{J}_{𝐦,𝐦+𝐞_x}^{S_x}+\widehat{J}_{𝐦𝐞_x,𝐦}^{S_x}`$ (44) $`+\widehat{J}_{𝐦,𝐦+𝐞_y}^{S_y}+\widehat{J}_{𝐦𝐞_y,𝐦}^{S_y}).`$ The non-zero term $`\widehat{F}_𝐦^{S_i}`$ on the righthand side of the spin continuity equation Eq. (41) is a formal expression, within the framework of bond spin current formalism, of the fact that spin is not conserved in SO coupled systems where it is forced into precession by the effective momentum-dependent magnetic field of the SO coupling. The fact that the bond spin current operator Eq. (33) appears in the spin continuity Eq. (41) as its divergence implies that its definition in Eq. (33) is plausible. However, the presence of the spin source operator $`\widehat{F}_𝐦^{S_i}`$ reminds us that such definition cannot be made unique, niu2005a unlike the case of the bond charge current which is uniquely determined by the charge continuity Eq. (19). If we evaluate the statistical average of Eq. (41) in a steady state (which can be either equilibrium or nonequilibrium), we obtain the identity $`{\displaystyle \underset{k=x,y}{}}\left(\widehat{J}_{𝐦,𝐦+𝐞_k}^{S_i}\widehat{J}_{𝐦𝐞_k,𝐦}^{S_i}\right)=\widehat{F}_𝐦^{S_i}.`$ (45) In particular, for the spin-$`S_z`$ component we get $`{\displaystyle \underset{k=x,y}{}}\left(\widehat{J}_{𝐦,𝐦+𝐞_k}^{S_z}\widehat{J}_{𝐦𝐞_k,𝐦}^{S_z}\right)`$ $`={\displaystyle \frac{t_{\mathrm{SO}}}{t_\mathrm{o}}}{\displaystyle \underset{k=x,y}{}}\left(\widehat{J}_{𝐦,𝐦+𝐞_k}^{S_k}+\widehat{J}_{𝐦𝐞_k,𝐦}^{S_k}\right),`$ (46) which relates the divergence of the spin-$`S_z`$ current (lefthand side) to the spin-source (righthand side) determined by the sum of the longitudinal component of the spin-$`S_x`$ current and the transverse component of the spin-$`S_y`$ current. Since no experiment has been proposed to measure local spin current density within the SO coupled sample, defined through Eq. (7) or its lattice equivalent Eq. (36), we can obtain additional well-defined and measurable information about the spin fluxes within the sample by computing the local spin density $`\widehat{𝐒}_𝐦`$ $`=`$ $`{\displaystyle \frac{\mathrm{}}{2}}{\displaystyle \underset{\alpha ,\beta =,}{}}𝝈_{\alpha \beta }\widehat{c}_{𝐦\alpha }^{}\widehat{c}_{𝐦\beta }`$ (47) $`=`$ $`{\displaystyle \frac{\mathrm{}}{4\pi i}}{\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑E{\displaystyle \underset{\alpha ,\beta =,}{}}𝝈_{\alpha \beta }G_{\mathrm{𝐦𝐦},\alpha \beta }^<(E)`$ $`=`$ $`{\displaystyle \frac{\mathrm{}}{4\pi i}}{\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑E\mathrm{Tr}_\mathrm{s}\left\{𝝈𝐆_{\mathrm{𝐦𝐦}}^<(E)\right\}.`$ Plotting of this quantity for Rashba SO coupled systems will provides us with insight into the stationary spin flow in the nonequilibrium steady transport state. ### I.3 Spin-resolved Landauer-Keldysh Green functions for finite-size mesoscopic devices The formalism discussed thus far does not depend on the details of the external driving force which pushes the system into a nonequilibrium state. That is, the system can be driven by either the homogeneous electric field applied to an infinite homogeneous 2DEG or the voltage (i.e., electrochemical potential) difference between the electrodes attached to a finite-size mesoscopic sample. For example, in the latter case, the external bias voltage only shifts the relative chemical potentials of the reservoirs into which the longitudinal leads (employed to simplify the boundary conditions) eventually terminate, so that the electrons do not feel any electric field in the course of ballistic propagation through clean 2DEG central region. The information about these different situations is encoded in the lesser Green function $`𝐆^<`$. Here we focus on experimentally accessible spin Hall bridges nikolic\_mesoshe where finite-size central region (C), defined on the $`L\times L`$ lattice, is attached to four external semi-infinite leads of the same width $`L`$. The leads at infinity terminate into the reservoirs where electrons are brought into thermal equilibrium, characterized by the Fermi-Dirac distribution function $`f(EeV_p)`$, to ensure the steady-state transport—in such Landauer setup baranger1989a current is limited by quantum transmission through a potential profile while power is dissipated non-locally in the reservoirs. The voltage in each lead of the four-terminal spin Hall bridge is $`V_p`$ ($`p=1,\mathrm{},4`$) so that the on-site potential $`\epsilon _𝐦`$ within the leads has to be shifted by $`eV_p`$. The spin-dependent lesser Green function $`𝐆^<`$ defined in Eq. (15) is evaluated within the finite-size sample region as a $`2L^2\times 2L^2`$ matrix in the site$``$spin space through the spin-resolved Keldysh equation for matrices nikolic\_accumulation $`𝐆^<(E)`$ $`=`$ $`𝐆(E)𝚺^<(E)𝐆^{}(E),`$ (48) which is valid in this form for steady-state transport when transients have died away. keldysh1965a Within the effective single-particle picture, the retarded Green function can be obtained by inverting the Hamiltonian $$𝐆(E)=\left[E𝐈_C𝐇_CU_𝐦\underset{p}{}𝚺_p(EeV_p)\right]^1,$$ (49) where the self-energies $`𝚺^<(E)`$ $`=`$ $`i{\displaystyle \underset{p}{}}𝚪_p(EeV_p)f(EeV_p),`$ (50) $`𝚪_p(E)`$ $`=`$ $`i\left[𝚺_p(E)𝚺_p^{}(E)\right],`$ (51) $`𝚺_p(E)`$ $`=`$ $`𝐇_{pC}^{}\left[\left(E+i0_+\right)𝐈_p𝐇_p^{\mathrm{lead}}\right]^1𝐇_{pC},`$ (52) are exactly computable in the non-interacting electron approximation and without any inelastic processes taking place within the sample. They take into account the “interaction” of the SO coupled sample with the attached leads in the Landauer transport setup where they generate finite that electrons spends within the 2DEG before escaping through the leads toward the macroscopic thermalizing reservoirs. Here $`𝐈_C`$ is the $`2L^2\times 2L^2`$ identity matrix, while $`𝐈_p`$ is the identity matrix in the infinite site$``$spin space of the $`p`$th lead, and we use the following Hamiltonian matrices $`\left(𝐇_C\right)_{\mathrm{𝐦𝐦}^{},\sigma \sigma ^{}}`$ $`=`$ $`1_{𝐦\sigma }\left|\widehat{H}\right|1_{𝐦^{}\sigma ^{}},(𝐦,𝐦^{}C),`$ $`\left(𝐇_p^{\mathrm{lead}}\right)_{\mathrm{𝐦𝐦}^{},\sigma \sigma ^{}}`$ $`=`$ $`1_{𝐦\sigma }\left|\widehat{H}\right|1_{𝐦^{}\sigma ^{}},(𝐦,𝐦^{}p),`$ $`\left(𝐇_{pC}\right)_{\mathrm{𝐦𝐦}^{},\sigma \sigma ^{}}`$ $`=`$ $`1_{𝐦\sigma }\left|\widehat{H}\right|1_{𝐦^{}\sigma ^{}},(𝐦p,𝐦^{}\text{C}),`$ with $`|1_{𝐦\sigma }`$ being a vector in the Fock space (meaning that the occupation number is one for the single particle state $`|𝐦\sigma `$ and zero otherwise) and $`\widehat{H}`$ is the Hamiltonian given in Eq. (10). In the general case of arbitrary applied bias voltage, the gauge invariance of measurable quantities (such as the current-voltage characteristic) with respect to the shift of electric potential everywhere by a constant $`V`$, $`V_pV_p+V`$ and $`U_𝐦U_𝐦+V`$, is satisfied on the proviso that the retarded self-energies $`𝚺_p(EeV_p)`$ introduced by each lead depend explicitly on the applied voltages at the sample boundary, while the computation of the retarded Green function $`𝐆(E)`$ has to include the electric potential landscape $`U_𝐦`$ within the sample christensen1996a \[which can be obtained from the Poisson equation with charge density Eq. (32)\]. However, when the applied bias is low, so that linear response zero-temperature quantum transport takes place through the sample \[as determined by $`𝐆(E_F)`$\], the exact profile of the internal potential becomes irrelevant. baranger1989a ; nikolic1999a ## II Spatial distribution of local spin currents and spin densities in ballistic four-terminal Rashba SO coupled nanostructures Under the time reversal transformation $`tt`$, the mass, charge, and energy are even while the velocity and Pauli matrices are odd. Thus, the charge currents are $`t`$-odd and must vanish in thermodynamic equilibrium (in the absence of magnetic field). On the other hand, the spin current-density operator Eq. (9), which is the product of two $`t`$-even quantities, can have its expectation values to be non-zero even in thermodynamic equilibrium. This has been explicitly demonstrated rashba2003a for its expectation values in the eigenstates of an infinite clean Rashba spin-split 2DEG. rashba2003a To investigate possible patterns of such equilibrium local spin currents in mesoscopic finite-size devices we plot in Fig. 1(a) the spatial distribution of the bond spin currents, carried by the whole Fermi sea, in a four-terminal ballistic device with no impurities where all leads are kept at the same potential. Although we find non-zero local spin currents, they do not transport any spin since the total spin current, obtained by summing the bond spin currents over an arbitrary transverse cross section of the device $$I_{\mathrm{trans}}^s(m_y)=\underset{m_x}{}\widehat{J}_{(m_x,m_y)(m_x,m_y+1)}^{S_z},$$ (54) or over any longitudinal cross section $$I_{\mathrm{long}}^s(m_x)=\underset{m_y}{}\widehat{J}_{(m_x,m_y)(m_x+1,m_y)}^{S_z},$$ (55) is identically equal to zero $`I_{\mathrm{trans}}^s(m_y)=I_{\mathrm{long}}^s(m_x)0`$. This picture also provides a direct microscopic proof that no equilibrium total spin currents, conjectured in Ref. pareek2004a, , can actually appear in the leads of an unbiased $`V_p=\mathrm{const}.`$ mesoscopic device in thermodynamic equilibrium, as demonstrated recently souma\_ringshe ; kiselev2004a within the Landauer-Büittker approach which operates only with the total spin and charge currents in the leads. In Fig.. 1(b) we apply low (i.e., linear response, see Sec. IV) bias voltage $`eV=10^3t_\mathrm{o}(E_FE_b)=0.2t_\mathrm{o}`$ between the longitudinal leads and integrate expression Eq. (37) from the bottom of the band to the chemical potential $`E_F+eV/2`$ of the left reservoir. In contrast to the equilibrium spin current density from Fig. 1(a), the vortex pattern is now distorted and non-zero total spin current $`I_{\mathrm{trans}}^s(m_y)0`$ in Eq. (54) emerges in the transverse direction, as expected in the phenomenology of the spin Hall effect. One of the highly unconventional features of the intrinsic spin Hall current is its dependence on the whole SO coupled Fermi sea, murakami2003a ; sinova2004a even when infinite system is driven out of equilibrium by the applied external electric field thereby limiting the charge transport and the extrinsic spin Hall response to the Fermi level through the nonequilibrium part of the distribution function. zhang2005a However, such property appears to be alien to the spirit of Fermi liquid theory where transport involves only quasiparticles with energies within $`k_BT`$ of the Fermi level. While the spin-currents crossing the transverse bonds in Fig. 1(b) are apparently carried by the whole Fermi sea, we now separate the integration in Eq. (37) for $`S_z`$ bond spin current into two parts $`\widehat{J}_{\mathrm{𝐦𝐦}^{}}^{S_z}`$ $`=`$ $`{\displaystyle \frac{t_\mathrm{o}}{2}}{\displaystyle \underset{E_b}{\overset{E_FeV/2}{}}}{\displaystyle \frac{dE}{2\pi }}\mathrm{Tr}\left[\widehat{\sigma }_z\left(𝐆_{\mathrm{𝐦𝐦}^{}}^<(E)𝐆_{𝐦^{}𝐦}^<(E)\right)\right]+{\displaystyle \frac{t_\mathrm{o}}{2}}{\displaystyle \underset{E_FeV/2}{\overset{E_F+eV/2}{}}}{\displaystyle \frac{dE}{2\pi }}\mathrm{Tr}\left[\widehat{\sigma }_z\left(𝐆_{\mathrm{𝐦𝐦}^{}}^<(E)𝐆_{𝐦^{}𝐦}^<(E)\right)\right]`$ (56) $`=`$ $`\widehat{J}_{\mathrm{𝐦𝐦}^{}}^{S_z(\mathrm{eq})}+\widehat{J}_{\mathrm{𝐦𝐦}^{}}^{S_z(\mathrm{neq})}.`$ The states from the band bottom $`E_b`$ to $`E_FeV/2`$ are fully occupied, while states in the energy interval from the electrochemical potential $`E_FeV/2`$ ($`e<0`$) of the right reservoir to the electrochemical potential $`E_F+eV/2`$ of the left reservoir are partially occupied because of the competition between the left reservoir which tries to fill them and the right reservoir which tries to deplete them. The profile of the first term $`\widehat{J}_{\mathrm{𝐦𝐦}^{}}^{S_z(\mathrm{eq})}`$ in Eq. (56) is shown in Fig. 1(c), while the spatial profile of the second term, representing the local spin current $`\widehat{J}_{\mathrm{𝐦𝐦}^{}}^{S_z(\mathrm{neq})}`$ carried by the states around the Fermi energy is shown in Fig. 1(d). The spatial distribution of the microscopic spin currents in Fig. 1(c) is akin to the vortex-like pattern of bond spin currents within the device in equilibrium in Fig. 1(a), and, therefore, does not transport any spin between two points in real space. Thus, Fig. 1 convincingly demonstrates that the non-zero spin Hall flux through the transverse cross sections in Fig. 1(b) and Fig. 1(d) is due to only the wave functions (or Green functions) at the Fermi energy (as $`T0`$), in accord with the general paradigms of the Landau’s Fermi liquid theory where transport quantities are expected to be expressed as the Fermi-surface property. baranger1989a ; haldane2004a We recall here that similar situation appears in charge transport in an external magnetic field where equilibrium (or persistent) current density, baranger1989a or bond charge currents in the lattice formalism, cresti2004a ; cresti2003a is non-zero even in unbiased devices (all leads at the same potential) in thermal equilibrium due to breaking of time-reversal invariance by the external magnetic field. However, such circulating or diamagnetic currents carried by the Fermi sea, which in Landauer-Keldysh formalism can be subtracted by separating the integration cresti2004a in a fashion similar to our Eq. (56), do not contribute to the net charge transport (i.e., to the total charge current measured in experiments) through any cross section of the device. baranger1989a ; cresti2004a Thus, early “Fermi sea” expressions for the linear response transport coefficients in, e.g., the quantum Hall effect theory baranger1989a or in the anomalous Hall effect theory, haldane2004a were eventually recast in terms of the Fermi surface determined quantities. Similarly, Fig. 1 demonstrates that spin Hall current carried by the “bulk” of the Fermi sea, which is equilibrium and does not really transport spin between two points in space, must be subtracted rashba2003a within any sensible theory for the intrinsic spin Hall conductivity defined in the thermodynamic limit of macroscopic systems. One of the basic tests for theories of the spin Hall effect is to predict the direction (i.e., the sign engel2005a ; nikolic\_accumulation ; niu2005a ) of the spin Hall current or the corresponding nonequilibrium spin Hall accumulation deposited by such current on the lateral boundaries of experimental devices. kato2004a ; wunderlich2005a The nonequilibrium (linear response) spin Hall current in Fig. 1(d) flows from the top to the bottom transverse lead because the spin-$``$ electrons are deflected to the right. This feature can be understood using the semiclassical picture based on the SO force operator nikolic\_soforce generated by the Rashba Hamiltonian Eq. (4) of the finite-size 2DEG $$\widehat{𝐅}_{\mathrm{SO}}=\frac{2\alpha ^2m^{}}{\mathrm{}^3}(\widehat{𝐩}\times 𝐳)\widehat{\sigma }^z\frac{dV_{\mathrm{conf}}(\widehat{y})}{d\widehat{y}}𝐲,$$ (57) and its expectation value in the spin-polarized wave packet state $`|\mathrm{\Psi }|`$. This apparently simple picture also explains why the transverse spin Hall current density bends toward the right in Fig. 1 while passing through the SO coupled region. However, this expectation value (i.e., the SO “force”) oscillates along the sample due to the precession of the deflected spin in the effective Rashba magnetic field which is nearly parallel to the $`y`$-axis because of the transverse confinement effects. nikolic\_soforce In ballistic strongly coupled SO structures such $`\alpha ^2`$-dependent SO “force”, which oscillates on the mesoscale set by the spin precession length $`L_{\mathrm{SO}}=\pi \mathrm{}^2/2m^{}\alpha =\pi t_\mathrm{o}a/2t_{\mathrm{SO}}`$ (on which the spin precesses by an angle $`\pi `$, i.e., the state $`|`$ evolves into $`|`$), will lead to a change of the sign nikolic\_mesoshe of spin Hall current as a function of the system size $`L/L_{\mathrm{SO}}`$. Also, since the mesoscopic spin Hall effect sensitively depends on the measurement geometry, nikolic\_mesoshe the sign of the transverse spin Hall current can change when non-ideal leads sheng2005a are attached to the sample. To highlight mesoscopic features baranger1989a (such as the effect of the measuring geometry and the properties of the attached probes) of the spin Hall effect in multiterminal ballistic SO coupled structure, nikolic\_mesoshe we plot in Fig. 2 the spatial profile of microscopic spin currents for the four-terminal bridge attached to the leads where Rashba SO coupling is switched on adiabatically (via a linear function) within the finite region of the leads adjacent to the 2DEG sample in the center of the device. In this measuring setup, the reflection nikolic\_purity at the interface separating zero and non-zero SO coupling regions is greatly suppressed, thereby enhancing the spin Hall current nikolic\_mesoshe \[as encoded by longer arrows in the profiles of Fig. 2 when compared to Fig. 1(d)\]. Since local spin current within the central region is not conserved, as manifested by the total spin current changing magnitude between different transverse cross sections (see Sec. IV for total spin current profiles) separated by long $`L_{\mathrm{SO}}`$, the spatial profiles of the local spin current density appear not to be directly measurable. rashba\_review However, the nonequilibrium spin density (i.e., the related spin magnetization) is a well-defined and measurable quantity. kato2004a ; wunderlich2005a Therefore, we plot the spatial distribution of the stationary flowing $`\widehat{S}_𝐦^z`$ Eq. (47) spin density in Fig. 2 to gain an additional insight into the dynamics of the spin Hall quantum transport. These microscopic picture convincingly demonstrate how spin-$``$ and spin-$``$ densities flow in opposite transverse directions through the attached leads, without any net charge flow in the transverse direction, thereby giving rise to a pure transverse spin Hall current. When we reverse the direction of the longitudinal charge current (by reversing the bias voltage $`VV`$), the transverse spin current and spin densities flip their sign, as exploited in experiments to confirm strong signatures of the spin Hall effect. wunderlich2005a Thus, in contrast to the arguments zhang2005a advocating impossibility of spin transport and accumulation by mechanisms driven solely by the intrinsic SO coupling terms in the effective Hamiltonian of spin-split semiconductors, Fig. 2 demonstrates that spin Hall effect originating in ballistic multi-terminal devices (without necessity for impurity induced effects) generates genuine nonequilibrium spin flux that can be used for spin injection and spintronics applications. zutic2004a Another mesoscale driven property of the spin Hall effect in ballistic nanostructures is its finite-size scaling being governed by the processes on the spin precession length $`L_{\mathrm{SO}}`$: One can differentiate the “mesoscopic” regime $`LL_{\mathrm{SO}}`$ where spin Hall current oscillates with increasing 2DEG size (changing sign with increasing size $`L\times L`$ of the 2DEG for strong enough Rashba SO coupling $`t_{\mathrm{SO}}0.04t_\mathrm{o}`$); and the “macroscopic” regime $`LL_{\mathrm{SO}}`$ where it saturates at some average value. nikolic\_mesoshe While the spatial profiles of local spin currents and spin densities are easy to interpret for $`L<L_{\mathrm{SO}}`$ (as in Fig. 1), Fig. 3 divulges how they become increasingly more intricate in the samples of the size $`L_{\mathrm{SO}}\times L_{\mathrm{SO}}`$ (for which the spin Hall current reaches its maximum nikolic\_mesoshe ), or in the macroscopic regime $`LL_{\mathrm{SO}}`$. Nevertheless, the total spin Hall current exists only in the transverse leads (the sum of bond $`S_z`$-spin currents over the cross section of longitudinal leads is equal to zero) with its magnitude and sign being identical to the terminal spin currents obtained within the Landauer-Büttiker formalism nikolic\_mesoshe (see Sec. IV for quantitative comparison). ## III Bulk vs. edge local spin currents in Disordered Four-Terminal Rashba coupled Nanostructures A surprising feature of the early predictions for the intrinsic spin Hall effect in hole-doped murakami2003a or electron-doped sinova2004a infinite homogeneous SO coupled semiconductor systems is apparent insensitivity of the spin Hall conductivity $`\sigma _{sH}`$ on the mean free path and relaxation rates. zhang2005a This has provoked intense theoretical scrutiny of the effects which spin-independent scattering off static impurities schliemann2004a imposes on the intrinsic spin Hall current, leading to the conclusion that, in fact, $`\sigma _{sH}0`$ vanishes inoue2004a ; sheng2005b for arbitrary small disorder in any model with SO coupling linear in momentum (such as the Rashba Hamiltonian of the spin-slit 2DEG) due to accidental cancellations. In the general case, where SO coupling contains higher order momentum terms, murakami2004a ; malshukov2005a $`\sigma _{sH}`$ can be resilient to sizable disorder strengths. chen2005a Moreover, an influential conjecture has emerged from the studies of the disorder effects on the intrinsic spin Hall current—macroscopic inhomogeneities facilitate spin currents rashba\_review ; shekhter2005a so that transverse edge spin current $`𝒥_y^z`$ could emerge near the sample-electrode interface, shytov2004a even in systems where it is expected to be destroyed in the bulk. inoue2004a However, the quantitative support for the picture of edge spin Hall currents is based on the analysis of semiclassical diffusive transport through a rather abstract structure, where 2DEG infinite in the transverse direction is attached to two massive electrodes in the longitudinal direction, shytov2004a that does not correspond to any experimentally realizable device. On the other hand, the presence of SO couplings makes the dynamics of transported spin in experimentally relevant confined structures strongly dependent on the properties of their interfaces, boundaries, and the attached electrodes, nikolic\_purity even for semiclassical spatial propagation of charges which carry spins evolving according to quantum dynamical laws. chang2004a For example, heuristic arguments based on the Keldysh formalism applied to an infinite two-terminal structure (lacking actual lateral edges) of Ref. shytov2004a, suggest that nonequilibrium spin Hall accumulation $`S_𝐦^z0`$ will appear only in the four corners at the lead/2DEG interfaces (due to $`𝒥_y^z0`$ existing within a spin relaxation length $`L_{\mathrm{SO}}`$ wide region around such interfaces), in contrast to the Keldysh formalism applied to finite-size 2DEG in the Landauer two-terminal setup where non-zero spin accumulation (with opposite sign on the two lateral edges kato2004a ; wunderlich2005a ) is found along the whole lateral edge. nikolic\_accumulation ; onoda2005a Within the formalism of bond spin currents of Sec. I (which represent the lattice version of the same quantity $`𝒥_y^z`$ studied in Ref. shytov2004a, ) these issues can be resolved through the exact evaluation of the retarded and lesser Green functions a non-interacting particle propagating through the random potential in finite-size multiterminal SO coupled device. Thus, we plot in Fig. 4 the spatial distribution of the local spin currents and spin densities for a single disordered Rashba SO coupled 2DEG, as well as their disorder-averages (which are to be compared with the analysis based on the diffusion equation shytov2004a ). The disorder strength is tuned to ensure the diffusive transport regime $`\mathrm{}L`$, while the magnitude of the spin Hall current in the leads at this concentration of impurities is still about $`80\%`$ of its maximum value set in the clean limit for the same four-terminal nanostructure. nikolic\_mesoshe We do not find any evidence for the confinement of spin fluxes near the 2DEG/longitudinal-leads interfaces. The conclusion based on the spatial profile of microscopic spin currents \[i.e., the “distribution of arrows” in Fig. 4(a)\] is further corroborated in Fig. 5 by the one-dimensional longitudinal profiles of the bond spin currents on different cross sections within the SO coupled sample and in the leads. Thus, both Fig. 4 and Fig. 5 suggest that precessing spins will propagate through the bulk of the diffusive Rashba SO coupled 2DEG in semiconductor heterostructure. ## IV Total spin Hall current vs. local spin Hall current: Landauer-Büttiker vs. Landauer-Keldysh picture The traditional charge transport experiments measure total current $`I`$ and the conductance $`I=GV`$ relating the charge current to the voltage drop $`V`$, rather than local current density $`𝐣`$ and the conductivity $`𝐣=\sigma 𝐄`$ relating it to the externally applied electric field (note also that in ballistic transport or quantum-coherent transport through a diffusive conductor conductivity ceases to exist as a local quantity baranger1989a ). Since realization of the total pure spin currents has been detected experimentally in optical pump-probe experiments, stevens2003 and several theoretical schemes are proposed to detect them indirectly via various electrical measurements, hirsch1999a ; hankiewicz2004a ; meier2003a we focus in this section on the properties of total spin Hall current $`I_{\mathrm{trans}}^s(m_y)`$ on different transverse cross sections of four-terminal devices, which is obtained from Eq. (54) by summing the bond spin currents. As shown in Fig. 6, the total pure spin current in the transverse leads, obtained by summing the nonequilibrium bond spin currents $`\widehat{J}_{\mathrm{𝐦𝐦}^{}}^{S_z(\mathrm{neq})}`$ over an arbitrary transverse cross section of the ideal leads (where SO coupling vanishes), flows through them in a conserved fashion, $`I_{\mathrm{trans}}^s(m_y)=\mathrm{const}.`$ for any $`m_y\mathrm{lead}`$. However, the same summation over the transverse cross sections within the 2DEG yields a quantity which is not conserved, except on the short length scales $`L_{\mathrm{SO}}`$. This is due to the fact that, e.g., injected eigenstate $`|`$ of $`\widehat{\sigma }_z`$ will precess in the effective magnetic field of the Rashba SO coupling (along the $`y`$-axis), thereby changing the amplitude of the spin current measured with respect to the $`z`$-axis as the spin quantization axis. Additional quantitative information about the microscopic details of spin fluxes is provided by Fig. 7 which plots the one-dimensional longitudinal profiles of the bond spin currents over the selected transverse cross sections (in the leads and in the 2DEG sample) cutting through the full spatial distributions of Fig. 1 and Fig. 3. Note that the sum of these longitudinal profiles yields the corresponding total spin current at the cross section $`m_y`$ in Fig. 6. The total spin currents in the leads in the linear response regime $`V0`$ can also be calculated using the spin-dependent Landauer-Büttiker scattering formalism for spin currents in multiprobe geometries. nikolic\_mesoshe ; souma\_ringshe ; pareek2004a In this formalism, one has to compute the spin-resolved transmission matrices connecting the spin-polarized asymptotic scattering states in semi-infinite ideal leads attached to the sample, nikolic\_purity which requires the knowledge of only the elements of the retarded Green function matrix Eq. (49) between the sites of the sample which are adjacent to the leads (similar expressions in terms of only the retarded Green function between the sample edges are obtained for the total charge currents caroli1971a ; nonoyama1998a ). We demonstrate in Fig. 8 that the spin Hall conductance obtained from the Landauer-Büttiker formalism $`G_{sH}^{z(\mathrm{LB})}=lim_{V0}I_2^s/V`$ (where 2 denotes the top transverse lead nikolic\_mesoshe ) is almost identical to the spin Hall conductance $`G_{sH}^{z(\mathrm{LK})}=I_{\mathrm{trans}}^s(m_y)/V`$ obtained by summing the bond spin currents over a cross section $`m_y`$ in the top transverse ideal lead, on the proviso that the applied bias voltage is small $`eVE_F`$ and current is carried only by the states at the Fermi level. This result further justifies the introduction of nonequilibrium bond spin current formula Eq. (37) for the $`z`$-component of spin. ## V Concluding Remarks In conclusion, we have demonstrated how to define the bond spin current, describing the spin flux across a single bond between two sites of the lattice model of an SO coupled semiconductors, and evaluate it in terms of the Keldysh nonequilibrium Green functions for to the Landauer setup where finite-size sample is attached to many semi-infinite ideal leads to form a theoretical model of experimentally accessible spin Hall bridges. Although spin current is not conserved within the SO coupled region (i.e., on the length scales comparable to the spin precession length $`L_{\mathrm{SO}}`$ on which the SO coupling manifests itself), the microscopic spin fluxes are nearly conserved on short scales. Thus, the bond spin currents make it possible to obtain their spatial distribution by following the dynamics of transported spin on the scale of the lattice spacing $`aL_{\mathrm{SO}}`$. Such profiles of (the lattice version of) local spin current density, together with stationary flow profiles of physically transparent (and measurable) local spin densities, allows us to demonstrate microscopic details of how pure transverse spin Hall current emerges in clean Rashba SO coupled 2DEG through which the unpolarized longitudinal charge current flows ballistically (where electrons do not scatter off impurities and do not feel the electric field). These spatial profiles are highly dependent on whether the 2DEG size is smaller or greater than the spin precession length, and can be affected by non-trivial measurement geometries, as discussed in the theory of the mesoscopic spin Hall effect. nikolic\_mesoshe The bond spin current within the bulk of the 2DEG is also resilient to weak disorder so that spin fluxes remain non-zero in the bulk of the sample and are not localized near the edges of the diffusive Rashba spin-split 2DEG. Using the bond spin current we explicitly demonstrate that nonequilibrium total spin current can be carried only by the states around the Fermi energy, while the Fermi sea contributes to local persistent spin currents which, however, do not transport any spin through a given cross section. The nanometer scale details of the spin Hall flow in multiterminal mesoscopic (quantum-coherent) structures convincingly show that the intrinsic to the crystal SO couplings can be used to generate spin fluxes, spin accumulation, and ultimately be employed to construct all-electrical spin injectors which do no require any ferromagnetic elements (whose coupling to semiconductors has been one of the major impediments for spintronics applications zutic2004a ). ###### Acknowledgements. We are grateful to J. Inoue and J. Sinova for valuable discussions. Acknowledgment is made to the donors of the American Chemical Society Petroleum Research Fund for partial support of this research.
warning/0506/quant-ph0506069.html
ar5iv
text
# Algebraic and information-theoretic conditions for operator quantum error-correction ## Abstract Operator quantum error-correction is a technique for robustly storing quantum information in the presence of noise. It generalizes the standard theory of quantum error-correction, and provides a unified framework for topics such as quantum error-correction, decoherence-free subspaces, and noiseless subsystems. This paper develops (a) easily applied algebraic and information-theoretic conditions which characterize when operator quantum error-correction is feasible; (b) a representation theorem for a class of noise processes which can be corrected using operator quantum error-correction; and (c) generalizations of the coherent information and quantum data processing inequality to the setting of operator quantum error-correction. To develop quantum technologies such as quantum computers and quantum communication networks, it will be necessary to protect quantum systems against the effects of noise. Considerable progress toward this goal was made in the late 1990s, when a theory of fault-tolerant quantum computing was developed Shor (1996); Kitaev (1997); Knill et al. (1998); Aharonov and Ben-Or (1997); Preskill (1998), based on the theory of quantum error-correcting codes Shor (1995); Steane (1996); Knill and Laflamme (1997); Gottesman (1996); Calderbank et al. (1998). The early theory of quantum error-correcting codes was based on the following ideas: (1) quantum information is stored in a subspace $`A`$ of a larger state space $`V=AC`$. $`A`$ is known as the *code space*, while $`V`$ is the state space of the physical system being used to store the information; (2) some physically-motivated noise process corrupts the physical system; (3) a recovery step is performed, restoring the original quantum information stored in $`A`$. Since its development this theory has been refined and generalized in a variety of ways, notably through the introduction of decoherence-free subspaces Palma et al. (1996); Duan and Duo (1997); Zanardi and Rasetti (1997); Lidar et al. (1998), noise-free subsystems Knill et al. (2000); Zanardi (2001); Kempe et al. (2001), and of operator quantum error-correction. In particular, the framework of *operator quantum error-correction* Kribs et al. (2005a, b) provides a single framework integrating and unifying all of these techniques. Operator quantum error-correction is based on the following ideas: (1) quantum information is stored in a space $`A`$ which appears as a tensor factor in a subspace of the overall state space, $`V`$, i.e., $`V=(AB)C`$; (2) some physically-motivated noise process corrupts the physical system; (3) a recovery step is performed, restoring the original encoded quantum information stored in $`A`$. Operator quantum error-correction is a significant generalization of standard quantum error-corection. Kribs *et al* Kribs et al. (2005a, b) have shown that operator quantum error-correction provides a natural framework unifying and generalizing earlier approaches, including standard quantum error-correction, decoherence-free subspaces, and noiseless subsystems. Bacon Bacon (2005) has recently exhibited interesting examples in which operator quantum error-correction plays a critical role. The purpose of this paper is to develop easily applied necessary and sufficient conditions for operator quantum error-correction. In particular, we obtain a set of algebraic conditions characterizing operator quantum error-correction. These conditions generalize the well-known conditions for standard quantum error-correction Bennett et al. (1996); Knill and Laflamme (1997), which are the basis for the theory of quantum error-correcting codes, enabling the construction of large classes of codes Gottesman (1996); Calderbank et al. (1997). The necessity of these conditions for operator quantum error-correction was proved in Kribs et al. (2005a), but the proof of sufficiency was left open. We establish the sufficiency of these conditions, and use the conditions to establish an elegant representation theorem for a class of noise processes which can be corrected using operator quantum error-correction. We also prove a set of information-theoretic conditions characterizing operator quantum error-correction, based on generalizations of the coherent information and the quantum data processing inequality. In the context of quantum error-correction codes these concepts were developed in Schumacher and Nielsen (1996), and were critical in developing the theory of quantum channel capacity Schumacher (1996); Schumacher and Nielsen (1996); Barnum et al. (1998a, b); Lloyd (1997); Shor (2002); Devetak (2005). Definition of operator quantum error-correction: Suppose $`V`$ is the Hilbert space for some quantum system, and we decompose $`V=(AB)C`$ for some choice of $`A,B`$ and $`C`$. Suppose $``$ is a quantum operation acting on $`V`$. Then we say $`A`$ is an $``$-correcting subsystem with respect to the decomposition $`V=(AB)C`$ if there exists a trace-preserving quantum operation $``$ (the recovery operation) such that for all $`\rho `$ with support on $`A`$, and all $`\sigma `$ with support on $`B`$, we have $`()(\rho \sigma )\rho \sigma ^{}`$, for some $`\sigma ^{}`$ with support on $`B`$. Physically, this means that we can store information in the subsystem $`A`$, and recover the information after noise $``$ by applying the recovery operation $``$. Quantum error-correcting codes arise as the special case of this definition where $`B`$ is trivial (i.e., one-dimensional), which is equivalent to decomposing $`V=AC`$. That is, in an error-correcting code we encode information in a *subspace*, while in an error-correcting subsystem we may encode information in a *subsystem* of a subspace. Algebraic characterization of operator quantum error-correction: Suppose $`(\rho )=_jE_j\rho E_j^{}`$ is an operator-sum representation for $``$ in terms of operation elements $`E_j`$. We will prove that the following two conditions are equivalent: \[a\]: $`A`$ is an $``$-correcting subsystem with respect to the decomposition $`V=(AB)C`$. \[b\]: $`PE_j^{}E_kP=I_AB_{jk}`$ for all $`j`$ and $`k`$, where $`P`$ projects onto $`AB`$, and the $`B_{jk}`$ are operators on $`B`$. Condition \[b\] provides an elegant and easily-checkable set of necessary and sufficient conditions for operator quantum error-correction, generalizing the standard quantum error-correction conditions Bennett et al. (1996); Knill and Laflamme (1997). Proof that \[a\] implies \[b\]: This was proved in Kribs et al. (2005a), and is a straightforward generalization of the corresponding part of the proof of the quantum error-correction conditions as given in, e.g., Chap. 10 of Nielsen and Chuang (2000). One of the ideas used in the proof is used again later, so for completeness we give a brief outline. Suppose the recovery operation $``$ has operation elements $`R_j`$. Define an operation $`𝒫(\rho )P\rho P`$. Then it can be shown that $`𝒫=_A𝒩`$ for some operation $`𝒩`$ on system $`B`$. Standard results (see, e.g., Chap. 9 of Nielsen and Chuang (2000)) about the unitary freedom in operation elements imply that $`R_jE_kP=IN_{jk}`$ for some set of operators $`N_{jk}`$ acting on system $`B`$. Multiplying this equation by its adjoint, for a suitable choice of indices we obtain $`PE_l^{}R_j^{}R_jE_kP=IN_{jl}^{}N_{jk}`$. Summing over $`j`$ and using the fact that $``$ is trace-preserving (i.e., $`_jR_j^{}R_j=I`$) gives the result. QED We will give two proofs that \[b\] implies \[a\]. The first proof is deeper, and is based on a third equivalent condition, \[c\]; we prove \[b\] $``$ \[c\] $``$ \[a\]. \[c\] has many rich consequences, including the information-theoretic characterization of operator error-correction described later, and a beautiful representation theorem (described below) for correctable $``$ in the special case when $`V=AB`$. Our second proof that \[b\] implies \[a\] is a more straightforward extension of the standard quantum error-correction conditions. This proof is arguably simpler than the first, but does not appear to have the same rich consequences, and so we merely provide a sketch. To state condition \[c\] involves a somewhat elaborate construction involving auxiliary systems, inspired by Schumacher and Nielsen (1996). We introduce systems $`R_A`$ and $`R_B`$ whose Hilbert spaces are copies of $`A`$ and $`B`$, respectively. We define (unnormalized) maximally entangled states $`|\alpha _j|j|j`$ of $`R_AA`$ and $`|\beta _k|k|k`$ of $`R_BB`$. The state $`|\alpha |\beta `$ may be regarded as a joint state of $`R_AR_BV`$ in a natural way. Next, we introduce a system $`E`$ which will act as a model environment for the operation $``$. We suppose $`E`$ has an orthonormal basis $`|j`$ whose elements are in one-to-one correspondence with the operation elements $`E_j`$. Supposing $`|s`$ is some fixed initial state of $`E`$, we define a linear operation $`L`$ on $`VE`$ which has the action $`L|\psi |s_jE_j|\psi |j`$. Note that the effect of $`L`$ on $`VE`$, after tracing out, is equivalent to the action of $``$ on $`V`$. Define a state $`|\psi ^{}(I_{R_AR_B}L)|\alpha |\beta |s`$. $`|\psi ^{}`$ can be thought of as the combined state of $`R_AR_BVE`$ *after* the noise is applied. We define a corresponding density matrix $`\rho ^{}|\psi ^{}\psi ^{}|`$, and use notations like $`\rho _{R_BE}^{}`$ to denote the result when all systems but $`R_B`$ and $`E`$ are traced out. With these definitions we may state condition \[c\]. \[c\]: $`\rho _{R_AR_BE}^{}=\rho _{R_A}^{}\rho _{R_BE}^{}`$. Proof that \[b\] implies \[c\]: The definition of $`\rho ^{}`$ and a direct calculation shows that: $`\rho _{R_AR_BE}^{}={\displaystyle \underset{jk}{}}PE_j^TE_k^{}P|jk|,`$ (1) where $`PE_j^TE_k^{}P`$ is understood as an operator on $`R_AR_B`$. To do this we identify the bases $`|j_{R_A}`$ and $`|j_A`$, and take the complex conjugate and transpose with respect to this basis. Taking the complex conjugate of \[b\] and substituting gives the desired result. (The converse, that \[c\] implies \[b\], also follows directly from Eq. (1), although we will not need this implication.) QED Proof that \[c\] implies \[a\]: (c.f. Schumacher and Nielsen (1996)) We Schmidt decompose $`|\psi ^{}`$ with respect to the bipartite decomposition $`R_AR_BE:V`$. Making use of the fact that the Schmidt vectors of $`R_AR_BE`$ are eigenvectors of $`\rho _{R_AR_BE}^{}=\rho _{R_A}^{}\rho _{R_BE}^{}`$, this gives rise to the Schmidt form (this and subsequent states are only written up to normalization): $`|\psi ^{}={\displaystyle \underset{jk}{}}\sqrt{q_k}|j_{R_A}|k_{R_BE}|e_{jk}_V,`$ (2) where the $`|j_{R_A}`$ are orthonormal eigenvectors of $`\rho _{R_A}^{}`$, the $`|k_{R_BE}`$ and $`q_k`$ are orthonormal eigenvectors and eigenvalues of $`\rho _{R_BE}^{}`$, and the $`|e_{jk}_V`$ are orthonormal Schmidt vectors on $`V`$. Define an orthonormal set of projectors $`P_k_j|e_{jk}_Ve_{jk}|`$ acting on $`V`$. We define the first step of recovery $``$ to be performing a measurement of $`P_k`$, resulting in the state: $`|\psi _k^{}={\displaystyle \underset{j}{}}|j_{R_A}|k_{R_BE}|e_{jk}_V.`$ (3) The second and final step of recovery is to apply a unitary $`U_k`$ which takes $`|e_{jk}_V`$ to $`|j_A|s_B`$, where $`|s_B`$ is some standard state of $`B`$. The net effect of the recovery procedure is to produce the following state of $`R_AR_BVE`$: $`|\psi _k^{\prime \prime }={\displaystyle \underset{j}{}}|j_{R_A}|j_A|s_B|k_{R_BE}`$ (4) Thus, we have restored the initial maximal entanglement between $`R_A`$ and $`A`$. Summarizing, we have shown that if $`R_AA`$ and $`R_BB`$ each start out maximally entangled, and we apply the noise $``$ followed by the recovery $``$ to $`V`$, then the resulting state of $`R_AA`$ is the original maximally entangled state. Standard techniques (e.g., Schumacher (1996)) imply that we must have $`()(\rho \sigma )=\rho \sigma ^{}`$ for all $`\rho `$ on system $`A`$ and all $`\sigma `$ on system $`B`$. QED Representation theorem for correctable operations: When $`V=AB`$, i.e., when $`C`$ is trivial, the proof that \[c\] implies \[a\] has as a consequence the elegant representation $`=𝒰(_A𝒩_B)`$ for some noisy operation $`𝒩_B`$ on $`B`$ alone, and some unitary operation $`𝒰`$ on $`V`$. To see this, note that when $`V=AB`$ the recovery procedure may be modified, omitting the step where $`P_k`$ is measured, and instead simply applying a single unitary operation $`W|e_{jk}_V|j_A|k_B`$. If $`𝒲`$ is the quantum operation corresponding to $`W`$ then we see that $`𝒲=_A𝒩_B`$, so using $`𝒰𝒲^{}`$ gives the desired representation. QED Alternate proof that \[b\] implies \[a\] (sketch): Fix a state $`\sigma =|ss|`$ of $`B`$, and define a quantum operation $`_s(\rho )(\rho \sigma )`$ mapping states of $`A`$ to states of $`V`$. We will use condition \[b\] to show that there exists a *single* universal recovery operation $``$ which acts as a recovery operation for *all* $`_s`$. Linearity then implies that $`()(\rho \sigma )=\rho \sigma ^{}`$ for all $`\rho `$ and $`\sigma `$. To prove this, note that a set of operation elements for $`_s`$ is the set $`E_{j,s}:AV`$ defined by $`E_{j,s}E_jP|s`$. That is, $`_s(\rho )=_jE_{j,s}\rho E_{j,s}^{}`$. (This can be verified by a calculation.) We will show that the set of errors $`E_{j,s}`$, where $`j`$ and $`|s`$ are *both* allowed to vary over all possible values, is a correctable set of errors mapping $`A`$ to $`V`$, in the sense of standard error-correction. This suffices to establish the existence of a single universal recovery operation $``$ which acts as a recovery operation for all $`_s`$. To see this, note that using \[b\] we obtain $`I_AE_{j,s}^{}E_{k,t}I_A=s|PE_j^{}E_kP|t=e_{jkst}I_A,`$ (5) for complex numbers $`e_{jkst}`$. Thus the standard error-correction conditions apply, which suffices to establish the existence of a suitable recovery $``$. QED Linearity of the set of correctable errors: Physically, one of the most important facts about quantum error-correction is that if $``$ is a recovery operation for a quantum operation $``$ with operation elements $`E_k`$, then $``$ also acts as a recovery operation for any quantum operation $``$ whose operation elements $`F_l`$ can be expressed as linear combinations of the $`E_k`$. It is this fact which allows us to focus attention on correcting a discrete set of errors (usually the Pauli $`I,X,Y`$ and $`Z`$ errors) since an arbitrary operation element on a qubit may be expressed as a linear combination of those errors. The analogous fact is also true for operator quantum error-correction. Suppose $``$ is a recovery operation for $``$, with respect to the decomposition $`V=(AB)C`$. As noted in the proof that \[a\] implies \[b\], we have $`R_jE_kP=IN_{jk}`$ for some set of operators $`N_{jk}`$ on $`B`$. Suppose $``$ is some other quantum operation whose operation elements $`F_l`$ may be expressed as linear combinations of the $`E_k`$, i.e., $`F_l=_ke_{lk}E_k`$, where the $`e_{lk}`$ are complex numbers. Then it follows that $`R_jF_lP=I\stackrel{~}{N}_{jl}`$, where $`N_{jl}_ke_{lk}N_{jk}`$. A direct computation shows that $``$ also acts as a recovery operation for $``$, which concludes the proof. QED Generalizations of operator quantum error-correction? We have studied the storage of quantum information in a subsystem $`A`$ of a subspace of $`V=(AB)C`$. Is it possible to store quantum information in some other way within $`V`$? For example, perhaps it is possible to decompose $`A`$ into two subspaces, $`A=A_1A_2`$, and store information solely in $`A_1`$. However, if we do this then the total vector space may be decomposed as $`V=(A_1B)\stackrel{~}{C}`$, where $`\stackrel{~}{C}=(A_2B)C`$, and thus this is a special case of the type of decomposition already considered. More generally, the distributive properties of the tensor product and direct sum ensure that no matter how we try to “nest” information within multiple layers of subspaces and subsystems, the end result can always be expressed as a decomposition of the form $`V=(AB)C`$, where the subsystem $`A`$ is used to store the quantum information. Information-theoretic characterization of correctability: For quantum error-correcting codes an information-theoretic necessary and sufficient condition for the correctability of trace-preserving $``$ was found in Schumacher and Nielsen (1996), and subsequently generalized to non-trace-preserving $``$ in Nielsen et al. (1998). We now find a set of information-theoretic necessary and sufficient conditions for operator quantum error-correction, generalizing the earlier conditions, and actually simplifying those in Nielsen et al. (1998). Most of the work has already been done in arriving at condition \[c\], above. Suppose we normalize the state $`|\psi ^{}`$ so $`\rho ^{}`$ and the corresponding reduced density matrices all have trace 1. The subadditivity inequality for entropy (see p. 515 and 516 of Nielsen and Chuang (2000)) implies that $`S(\rho _{R_AR_BE}^{})S(\rho _{R_A}^{})+S(\rho _{R_BE}^{})`$, with equality if and only if $`\rho _{R_AR_BE}^{}=\rho _{R_A}^{}\rho _{R_BE}^{}`$. It follows that a necessary and sufficient condition for $``$ to be correctable is that $`S(\rho _{R_A}^{})+S(\rho _{R_BE}^{})=S(\rho _{R_AR_BE}^{}`$. This may be rewritten in a more convenient form by noting that $`S(\rho _{R_A}^{})=S(\rho _{R_A})=S(\rho _A)`$, and that $`S(\rho _{R_AR_BE}^{})=S(\rho _V^{})`$. This gives us the following necessary and sufficient condition for $``$ to be correctable. (Note that in an obvious notation $`S(\rho _A)=\mathrm{log}(d_A)`$, where $`d_A`$ is the dimension of system $`A`$, since $`A`$ is initially maximally entangled with $`R_A`$.) \[d\]: $`S(\rho _A)=S(\rho _V^{})S(\rho _{R_BE}^{})`$. The conditions \[d\] generalize the necessary and sufficient conditions in Schumacher and Nielsen (1996); Nielsen et al. (1998) (c.f. Schumacher and Westmoreland (2002); Ogawa (2005)), which correspond to the case when $`B`$ is trivial. Note that Schumacher and Nielsen (1996); Nielsen et al. (1998) allow $`A`$ and $`R_A`$ to start out in a state which is not maximally entangled, but rather are merely of full Schmidt rank. Our arguments are easily generalized to this case. Data processing inequality: We have described the condition \[d\] as information-theoretic, but have not suggested an information-theoretic interpretation of the quantities involved. Such an interpretation is suggested by the following argument, which generalizes the coherent information introduced in Schumacher and Nielsen (1996). Schumacher and Nielsen (1996) showed that the coherent information satisfied a monotonicity property known as the *quantum data processing inquality*, which states that quantum information can only ever be lost as it is passed through multiple quantum channels; once lost, quantum information can never be recovered. The coherent information and quantum data processing inequality played a key role in subsequent investigations of the quantum channel capacity Schumacher (1996); Schumacher and Nielsen (1996); Barnum et al. (1998a, b); Lloyd (1997); Shor (2002); Devetak (2005). We now prove an analogue of the quantum data processing inequality which applies to operator quantum error-correction. Our analysis is based on the conditional entropy of $`R_A`$ given $`V`$, $`S(R_A|V)S(V)S(R_AV)`$, which generalizes the coherent information. The following argument suggests that this may be regarded as a measure of the amount of quantum information about the initial state of $`A`$ which is still stored in $`V`$. Suppose we apply a sequence of trace-preserving quantum operations $`_1,_2,\mathrm{}`$ to $`V`$. Standard monotonicity properties of the conditional entropy imply that $`S(R_A|V)S(R_A^{}|V^{})S(R_A^{\prime \prime }|V^{\prime \prime })\mathrm{},`$ (6) where a single prime indicates that $`_1`$ has been applied, a double prime indicates that $`_2_1`$ has been applied, and so on. Eq. (6) is a generalization of the data processing inequality obtained in Schumacher and Nielsen (1996). Condition \[d\] is easily seen to be equivalent to the condition $`S(R_A^{}|V^{})=S(R_A|V)`$, i.e., that the coherent information be preserved by the operation $``$. Indeed, a consequence of (6) is an informative alternative proof of the necessity of \[d\]. Suppose $`_1=`$ and $`_2=`$. The fact that $``$ restores the information stored in $`A`$ implies that $`S(R_A|V)=S(R_A^{\prime \prime }|V^{\prime \prime })`$. It follows from (6) that we must have $`S(R_A^{}|V^{})=S(R_A|V)`$, which implies \[d\]. Conclusion: Operator quantum error-correction is a recently introduced technique for stabilizing quantum information, which generalizes and unifies previous approaches, including standard quantum error-correcting codes, decoherence-free subspaces, and noiseless subsystems. In this paper we’ve developed algebraic and information-theoretic necessary and sufficient conditions for operator quantum error-correction, and used these conditions to develop an elegant representation theorem for a wide class of correctable noise processes, as well as generalizations of the coherent information and quantum data processing inequality. Open problems include the systematic investigation of specific operator quantum codes, and the investigation of techniques for fault-tolerant quantum information processing using operator quantum codes. ###### Acknowledgements. Thanks to Dave Bacon, Steve Bartlett, Dominic Berry, Jennifer Dodd, and Andrew Doherty for helpful discussions and suggestions for improvement.
warning/0506/cond-mat0506761.html
ar5iv
text
# Rashba spin-orbit coupling and spin precession in carbon nanotubes ## 1 Introduction Spintronics in molecular conductors is a field attracting more and more attention, both from fundamental physics as well as from application-oriented material science . Here the quantum-mechanical electronic spin is the central object controlling transport properties. For a conductor sandwiched between ferromagnetic leads, a different resistance can be observed depending on the relative orientation of the lead magnetizations. Quite often, the resistance is larger in the antiparallel configuration than in the parallel one, but sometimes also the reverse situation can be observed. It is useful to define the tunnel magnetoresistance (TMR), $`\rho _t=(R_{AP}R_P)/R_P`$, as the relative difference between the corresponding resistances. A particularly interesting material in that context is provided by carbon nanotubes (CNTs), see Refs. for general reviews. Quite a number of experimental studies concerning spin transport through individual multi- (MWNT) or single-walled (SWNT) nanotubes contacted by ferromagnetic leads have been reported over the past few years . In particular, the experiments of the Basel group use thin-film PdNi alloys as ferromagnetic leads in order to contact either SWNTs or MWNTs, where the shape anisotropy and the geometry of the setup allow for the study of the spin-dependence of electrical transport. These experiments have revealed oscillatory behavior of the TMR as a function of the external gate voltage. Similar oscillations were predicted as a consequence of the gate-voltage-tunable Rashba spin-orbit (SO) interaction in a classic paper by Datta and Das some time ago . Since Datta-Das oscillations have still not been observed experimentally so far, a thorough theoretical investigation of this effect in nanotubes is called for and provided here. Unfortunately, from our analysis below, we find that the weakness of SO couplings in nanotubes excludes an interpretation of these data in terms of the Datta-Das effect – they can, however, be explained in terms of quantum interference effects . Nevertheless, we show that the presence of multiple bands in CNTs is not detrimental, and under certain circumstances, the effect may be sufficiently enhanced to be observable, e.g., by a tuning of the number of bands via external gates along the lines of Ref. . In the original Datta-Das proposal , subband mixing was ignored so that different channels just add up coherently, but subband mixing has later been argued to spoil the effect . In CNTs, the special band structure requires a careful re-examination of the Datta-Das idea in this context, and we shall show that the arguments of Refs. do not necessarily apply here. Recent theoretical studies of spin-dependent transport in CNTs have mainly focused on the single-channel limit, taking into account electron-electron interactions within the framework of the Luttinger liquid theory (see also for related discussions on interacting quantum wires with Rashba SO coupling). Here we confine ourselves to the noninteracting problem in order to not overly complicate the analysis, but study the many-band case and details of the band structure. Interactions can be taken into account within the Luttinger liquid approach at a later stage, and may enhance the effect of SO couplings . We shall also neglect disorder effects. Mean free paths in high-quality SWNTs typically exceed $`1\mu `$m, while in MWNTs this may be a more severe approximation for some samples. However, high-quality MWNTs with ultra-long mean free paths have also been reported recently . The structure of this paper is as follows. In Sec. 2 we derive the Rashba spin-orbit hamiltonian from microscopic considerations. The resulting tight-binding SO hamiltonian will be studied at low energy scales in Sec. 3, where we derive its continuum form. In Sec. 4, the consequences with regard to Datta-Das oscillations in the TMR are analyzed. We shall always consider the zero-temperature limit, and (in most of the paper) put $`\mathrm{}=1`$. ## 2 Rashba spin-orbit coupling in nanotubes We start by noting that transport effectively proceeds through the outermost shell of a MWNT only, such that we can take a single-shell model even when dealing with a MWNT. Experimentally and theoretically, it is understood that such a model works very well in good-quality MWNTs , essentially because only the outermost shell is electrically contacted and tunneling between different shells is largely suppressed . Naturally, a single-shell description is also appropriate for SWNTs, where we assume a sufficiently large radius $`R`$ such that occupation of multiple subbands can be possible. (For a MWNT, $`R`$ denotes the radius of the outermost shell.) Depending on the electrochemical potential $`\mu `$ (doping level), we then have to deal with $`N`$ spin-degenerate bands. We assume full quantum coherence (no dephasing), so that the usual Landauer-Büttiker approach applies, and exclude external magnetic fields or electric field inhomogeneities, say, due to the electrodes. We proceed to derive the Rashba SO interaction, $`H_{so}`$, for this problem. Notice that this is different from the intrinsic atomic SO interaction discussed in Refs. . In particular, the SO coupling in Refs. vanishes in the limit of large radius, which is not the case for the Rashba SO coupling we discuss below. Though Ando’s SO coupling could straightforwardly be included in our analysis, being gate-voltage independent it could not change our conclusions relative to the gate-voltage dependent oscillations in the magnetoresistance and is neglected in what follows. We first define a fixed reference frame $`𝒮=\{\widehat{Y},\widehat{Z},\widehat{X}\}`$, with unit vector $`\widehat{X}`$ pointing in the axis direction and $`\widehat{Z}`$ perpendicular to the substrate on which the CNT is supposed to be located. Next we introduce a second, local reference frame $`𝒮_i=\{\widehat{\rho }_i,\widehat{t}_i,\widehat{X}\}`$ relative to each lattice site $`\stackrel{}{R}_i`$ on the tube surface, where $`\widehat{\rho }_i`$ and $`\widehat{t}_i`$ are unit vectors along the local normal and tangential (around the circumference) directions at $`\stackrel{}{R}_i`$, respectively. Using polar coordinates in the plane transverse to the tube axis, the relation between $`𝒮`$ and $`𝒮_i`$ is given by $$\widehat{\rho }_i=\mathrm{cos}\phi _i\widehat{Y}+\mathrm{sin}\phi _i\widehat{Z},\widehat{t}_i=\mathrm{sin}\phi _i\widehat{Y}+\mathrm{cos}\phi _i\widehat{Z}.$$ (1) The position vector of a given carbon atom can then be written as $`\stackrel{}{R}_i=R\widehat{\rho }_i+X_i\widehat{X}`$. For later convenience, we introduce also another reference frame. For each pair of sites $`\stackrel{}{R}_i`$ and $`\stackrel{}{R}_j`$, we define $$\stackrel{}{R}_{ij}=\stackrel{}{R}_i\stackrel{}{R}_jX_{ij}\widehat{X}+\stackrel{}{\rho }_{ij},$$ (2) and denote the direction perpendicular to $`\widehat{\rho }_{ij}`$ and $`\widehat{X}`$ as $`\widehat{\rho }_{ij}^{}`$. Then $`\{\widehat{\rho }_{ij}^{},\widehat{\rho }_{ij},\widehat{X}\}`$ constitutes a new local frame $`𝒮_{ij}`$, and one has $`\widehat{\rho }_{ij}^{}`$ $`=`$ $`\mathrm{cos}[(\phi _i+\phi _j)/2]\widehat{Y}+\mathrm{sin}[(\phi _i+\phi _j)/2]\widehat{Z},`$ (3) $`\widehat{\rho }_{ij}`$ $`=`$ $`\mathrm{sin}[(\phi _i+\phi _j)/2]\widehat{Y}+\mathrm{cos}[(\phi _i+\phi _j)/2]\widehat{Z}.`$ The $`2p_z`$ orbital at position $`\stackrel{}{R}_i`$ can then be represented as $$\chi _i(\stackrel{}{r}\stackrel{}{R}_i)=\alpha (\stackrel{}{r}\stackrel{}{R}_i)\widehat{\rho }_ie^{\beta |\stackrel{}{r}\stackrel{}{R}_i|},$$ (4) where $`4\alpha =(2\pi a_0^5)^{1/2}`$, $`\beta =(2a_0)^1`$, $`a_0=\mathrm{}^2/me^2=0.53`$Å is the Bohr radius, and $`m`$ is the electron’s mass. We introduce an index $`i`$ on the orbital in order to keep track of the atom at which it is centered. The wavefunction (4) is expected to be highly accurate for not too small $`R`$, where hybridization with the $`sp^2`$ orbitals is negligible. At large distances from the tube, external gates generally produce an electric field perpendicular to the tube axis and the substrate. As it has been shown in detail in previous works , polarization effects of the CNT itself due to a transverse field result in a reduction of the externally applied field described by $$E_0=\frac{1}{1+2\alpha _{0yy}/R^2}E_{ext},$$ where $`\alpha _{0yy}`$ is the unscreened transverse static polarizability. Since $`\alpha _{0yy}`$ is approximately proportional to $`R^2`$, the factor in front of $`E_{ext}`$ practically equals a constant, $`0.2`$ . Then, assuming homogeneity, the electric field due to the gate can be written as $$\stackrel{}{E}=E_0\widehat{Z},$$ (5) which in turn produces the (first-quantized) Rashba spin-orbit interaction . With standard Pauli matrices $`\stackrel{}{\sigma }`$ acting in spin space, $$H_{so}=\frac{e\mathrm{}}{4m^2c^2}\stackrel{}{E}(\stackrel{}{\sigma }\times \stackrel{}{p}).$$ (6) We proceed to derive the second-quantized spin-orbit hamiltonian within the tight-binding approximation. For that purpose, we need the matrix element of the momentum operator between two $`2p_z`$ orbitals $`\stackrel{}{p}_{ij}=\chi _i|\stackrel{}{p}|\chi _j`$, from which we get the following form for the SO hamiltonian: $$H_{so}=g\underset{ij}{}c_i^{}\left[(\stackrel{}{\sigma }\times \stackrel{}{p}_{ij})\widehat{Z}\right]c_j,$$ (7) where the fermionic operator $`c_{i\sigma }`$ destroys an electron with spin $`\sigma =,`$ in the $`2p_z`$ orbital centered at $`\stackrel{}{R}_i`$, and $`g=E_0/4m^2c^2`$. For calculational convenience, the matrix element $`\stackrel{}{p}_{ij}`$ can be written as $`g\stackrel{}{p}_{ij}=i(\stackrel{}{v}_{ij}+\stackrel{}{u}_{ij})`$, where the spin-orbit vectors $`\stackrel{}{v}_{ij}`$ and $`\stackrel{}{u}_{ij}`$ are defined as $`\stackrel{}{v}_{ij}`$ $`=`$ $`g\alpha {\displaystyle d^3\stackrel{}{r}\chi _i(\stackrel{}{r}\stackrel{}{R}_i)\widehat{\rho }_je^{\beta |\stackrel{}{r}\stackrel{}{R}_j|}},`$ (8) $`\stackrel{}{u}_{ij}`$ $`=`$ $`g\beta {\displaystyle d^3\stackrel{}{r}\chi _i(\stackrel{}{r}\stackrel{}{R}_i)\frac{\stackrel{}{r}\stackrel{}{R}_j}{|\stackrel{}{r}\stackrel{}{R}_j|}\chi _j(\stackrel{}{r}\stackrel{}{R}_j)},`$ (9) Note that the modulus of $`\stackrel{}{v}_{ij}`$ and $`\stackrel{}{u}_{ij}`$ has dimension of energy, and their sum (but not necessarily each term separately) is antisymmetric under exchange of $`i`$ and $`j`$. We first observe that the spin-orbit vectors connecting a site with itself clearly vanish, since $`\chi _i|\stackrel{}{p}|\chi _i=0`$. Let us then discuss spin-orbit vectors connecting different sites. Since the orbitals (4) decay exponentially, it is sufficient to consider only the case of nearest neighbors. We start with $`\stackrel{}{v}_{ij}`$. Shifting $`\stackrel{}{r}\stackrel{}{s}+\stackrel{}{R}_i`$ in Eq. (8) and using Eq. (4), we obtain $$\stackrel{}{v}_{ij}=g\alpha ^2\widehat{\rho }_jd^3\stackrel{}{s}(\stackrel{}{s}\widehat{\rho }_i)e^{\beta s}e^{\beta |\stackrel{}{s}+\stackrel{}{R}_{ij}|}.$$ Using $`\stackrel{}{s}=s_{}\widehat{R}_{ij}+\stackrel{}{s}_{}`$, we then rewrite the above integral as $$d^3\stackrel{}{s}(s_{}\widehat{R}_{ij}+\stackrel{}{s}_{})\widehat{\rho }_ie^{\beta s}e^{\beta \sqrt{(s_{}+d)^2+s_{}^2}},$$ where we use $`|\stackrel{}{R}_{ij}|=d`$, with the nearest-neighbor distance among carbon atoms in graphene $`d=1.42`$ Å. Note that $`\beta d=1.34`$. The second term in the brackets is odd in $`\stackrel{}{s}_{}`$ and thus vanishes, and we obtain $$\stackrel{}{v}_{ij}=g\alpha ^2\widehat{\rho }_j\frac{2R}{d}\mathrm{sin}^2(\frac{\phi _i\phi _j}{2})d^4\gamma _0,$$ (10) where we have used $`\widehat{R}_{ij}\widehat{\rho }_i=\frac{2R}{d}\mathrm{sin}^2(\frac{\phi _i\phi _j}{2})`$ and the dimensionless numerical factor $`\gamma _0`$ : $$\gamma _0=𝑑x𝑑y𝑑zxe^{\beta d\sqrt{x^2+y^2+z^2}}e^{\beta d\sqrt{(x+d)^2+y^2+z^2}}.$$ For $`\stackrel{}{v}_{ji}`$, we find $$\stackrel{}{v}_{ji}=g\alpha ^2\widehat{\rho }_i\frac{2R}{d}\mathrm{sin}^2(\frac{\phi _i\phi _j}{2})d^4\gamma _0.$$ Notice that, up to higher orders in $`d/R`$, the unit vectors $`\widehat{\rho }_{i,j}`$ can be replaced by $`\widehat{\rho }_{ij}^{}`$, which makes clear that $`\stackrel{}{v}_{ij}`$ is normal to the tube surface. Now $`|\mathrm{sin}[(\phi _i\phi _j)/2]|`$ varies between zero (when the two sites are aligned in the axis direction) and $`d/2R1`$ (when the two sites are aligned in the circumferential direction). Thus, to zeroth order in $`d/R`$, $`\stackrel{}{v}_{ij}`$ vanishes: it is a pure curvature effect, peculiar of nanotubes, which does not exist in graphene. In practice, $`\stackrel{}{v}_{ij}`$ is tiny and certainly subleading to $`\stackrel{}{u}_{ij}`$, which turns out to be of order $`(d/R)^0`$. We shall therefore neglect it in what follows. Let us now turn to $`\stackrel{}{u}_{ij}`$. We shift $`\stackrel{}{r}\stackrel{}{s}+(\stackrel{}{R}_i+\stackrel{}{R}_j)/2`$ in Eq. (9), and rewrite $`\stackrel{}{u}_{ij}`$ as the sum of two terms: $`\stackrel{}{u}_{ij}^{(1)}`$ $`=`$ $`g\beta {\displaystyle d^3\stackrel{}{s}\chi _i(\stackrel{}{s}\stackrel{}{R}_{ij}/2)\chi _j(\stackrel{}{s}+\stackrel{}{R}_{ij}/2)\frac{\stackrel{}{s}}{|\stackrel{}{s}+\stackrel{}{R}_{ij}/2|}},`$ (11) $`\stackrel{}{u}_{ij}^{(2)}`$ $`=`$ $`{\displaystyle \frac{g\beta }{2}}\stackrel{}{R}_{ij}{\displaystyle d^3\stackrel{}{s}\chi _i(\stackrel{}{s}\stackrel{}{R}_{ij}/2)\chi _j(\stackrel{}{s}+\stackrel{}{R}_{ij}/2)\frac{1}{|\stackrel{}{s}+\stackrel{}{R}_{ij}/2|}}.`$ (12) Writing again $`\stackrel{}{s}=s_{}\widehat{R}_{ij}+\stackrel{}{s}_{}`$, the computation of the above integrals leads, to the lowest non-vanishing order in $`d/R`$, to the following expressions: $`\stackrel{}{u}_{ij}^{(1)}`$ $`=`$ $`g\beta \alpha ^2\stackrel{}{R}_{ij}d^4\gamma _1u_1\stackrel{}{R}_{ij},`$ (13) $`\stackrel{}{u}_{ij}^{(2)}`$ $`=`$ $`{\displaystyle \frac{g\beta }{2}}\alpha ^2\stackrel{}{R}_{ij}d^4\gamma _2u_2\stackrel{}{R}_{ij},`$ (14) with the dimensionless numerical factors $`\gamma _1`$ $`=`$ $`{\displaystyle 𝑑x𝑑y𝑑z\frac{xz^2e^{\beta d\sqrt{(x1/2)^2+y^2+z^2}}e^{\beta d\sqrt{(x+1/2)^2+y^2+z^2}}}{\sqrt{(x+1/2)^2+y^2+z^2}}}`$ $``$ $`0.0375,`$ and $`\gamma _2`$ $`=`$ $`{\displaystyle 𝑑x𝑑y𝑑z\frac{z^2e^{\beta d\sqrt{(x1/2)^2+y^2+z^2}}e^{\beta d\sqrt{(x+1/2)^2+y^2+z^2}}}{\sqrt{(x+1/2)^2+y^2+z^2}}}`$ $``$ $`0.3748.`$ To lowest order in $`d/R`$, it does not make a difference whether we take the tangent unit vector at $`\stackrel{}{R}_j`$, $`\stackrel{}{R}_i`$, or at $`(\stackrel{}{R}_i+\stackrel{}{R}_j)/2`$. Hence we may write $`\widehat{\rho }_{ij}\widehat{e}_\phi `$, where $`\widehat{e}_\phi `$ is the unit tangent vector at $`(\stackrel{}{R}_i+R_j)/2`$. We then get SO couplings along the axial and along the circumferential direction, $$\stackrel{}{u}_{ij}=u\left[(\stackrel{}{R}_{ij}\widehat{X})\widehat{X}+(\stackrel{}{R}_{ij}\widehat{e}_\phi )\widehat{e}_\phi \right],$$ (17) with $`u=u_1+u_2`$. Note that we have neglected a tiny component of $`\stackrel{}{R}_{ij}`$ normal to the tube surface. The above discussion then results in the tight-binding hamiltonian $`H=H_0+H_{so}`$, where $$H_0=t\underset{\stackrel{}{r},a}{}c_{B,\stackrel{}{r}+\stackrel{}{\delta }_a}^{}c_{A,\stackrel{}{r}}+h.c.,$$ with $`t2.7`$ eV . Here the $`\stackrel{}{r}`$ denote all sublattice-A tight-binding sites of the lattice. Furthermore, the $`\stackrel{}{\delta }_{a=1,2,3}`$ are vectors connecting $`\stackrel{}{r}`$ with the three nearest-neighbor sites which are all located on sublattice B . Since we consider the limit $`d/R1`$, the $`\stackrel{}{\delta }_a`$ at each site effectively lie in the tangent plane to the tube surface at that site. The Rashba spin-orbit hamiltonian then reads $$H_{so}=iu\underset{\stackrel{}{r},a}{}c_{B,\stackrel{}{r}+\stackrel{}{\delta }_a}^{}\left[\left(\stackrel{}{\sigma }\times [(\stackrel{}{\delta }_a\widehat{X})\widehat{X}+(\stackrel{}{\delta }_a\widehat{e}_\phi )\widehat{e}_\phi ]\right)\widehat{Z}\right]c_{A,\stackrel{}{r}}+h.c.$$ ## 3 Continuum limit Since we are interested in the low-energy long-wavelength properties, we now expand the electron operator around the Fermi points $`K,K^{}`$ in terms of Bloch waves , $$\frac{c_{p\stackrel{}{r}}}{\sqrt{S}}=e^{i\stackrel{}{K}\stackrel{}{r}}F_{1p}(\stackrel{}{r})+e^{i\stackrel{}{K}\stackrel{}{r}}F_{2p}(\stackrel{}{r}),$$ (18) where $`S=\sqrt{3}a^2/2`$ is the area of the unit cell, $`a=\sqrt{3}d`$, and $`p=A/B`$ is the sublattice index. The $`F_{\alpha p}`$ are slowly varying electron field operators, and we choose the Fermi points at $`\stackrel{}{K}=(4\pi /3a,0)`$ and $`\stackrel{}{K}^{}=\stackrel{}{K}`$ . We then expand $`F(\stackrel{}{r}+\stackrel{}{\delta })F(\stackrel{}{r})+\stackrel{}{\delta }F(\stackrel{}{r})`$ and use the bond vectors $$\stackrel{}{\delta }_1=\frac{a}{\sqrt{3}}(0,1),\stackrel{}{\delta }_2=\frac{a}{2}(1,1/\sqrt{3}),\stackrel{}{\delta }_3=\frac{a}{2}(1,1/\sqrt{3}).$$ (19) These vectors are given in a fixed reference frame for a 2D graphene sheet, and we then must perform a rotation to longitudinal and circumferential directions via the chiral angle. This rotation results in fixed phases that can be absorbed in the definition of $`F_{\alpha p}`$ and do not appear in final results. This is of course expected from the $`U(1)`$ symmetry emerging at low energies in the dispersion relation of graphene . After some algebra, the usual Dirac hamiltonian for the kinetic term follows, $$H_0=vd^2\stackrel{}{r}F^{}\left[(T_0\tau _2\sigma _0)(i_x)+(T_3\tau _1\sigma _0)(i_y)\right]F,$$ (20) where $`v=\sqrt{3}at/28\times 10^5`$ m/sec is the Fermi velocity, and $`x,y`$ are longitudinal and circumferential coordinates, respectively, with $`0<y2\pi R`$. Finally, $`T_i`$ and $`\tau _i`$ are also Pauli matrices that now act in the space of Fermi ($`K,K^{}`$) points and sublattice space ($`A,B`$), respectively. For $`i=0`$, these are defined as $`2\times 2`$ unit matrices. The low-energy limit of the SO term (2) can be obtained in the following way. First we observe that $$\left[(\stackrel{}{\delta }_a\widehat{X})\widehat{X}+(\stackrel{}{\delta }_a\widehat{e}_\phi )\widehat{e}_\phi \right]\times \widehat{Z}=(\stackrel{}{\delta }_a\widehat{X})\widehat{Y}\mathrm{sin}(y/R)(\stackrel{}{\delta }_a\widehat{e}_\phi )\widehat{X}.$$ Here the only approximation is the assumption that the bond vectors lie in the plane tangent to the nanotube surface at $`\stackrel{}{r}`$. Second, by using the bond vectors (19) and taking into account the chiral angle $`\eta `$ between the fixed direction on the graphite sheet and the circumferential direction on the nanotube, one obtains $`{\displaystyle \underset{a}{}}c_{\stackrel{}{r}+\stackrel{}{\delta }_a}^{}\left(\stackrel{}{\delta }_a\widehat{X}\right)c_\stackrel{}{r}`$ $``$ $`{\displaystyle \frac{3d}{2}}\left(F_{B1}^{}e^{i\eta }F_{A1}+F_{B2}^{}e^{i\eta }F_{A2}\right),`$ $`{\displaystyle \underset{a}{}}c_{\stackrel{}{r}+\stackrel{}{\delta }_a}^{}\left(\stackrel{}{\delta }_a\widehat{e}_\phi \right)c_\stackrel{}{r}`$ $``$ $`{\displaystyle \frac{3d}{2}}\left(iF_{B1}^{}e^{i\eta }F_{A1}iF_{B2}^{}e^{i\eta }F_{A2}\right).`$ Notice that we take into account exactly the relative orientation of the bond vectors with respect to the directions $`\widehat{X}`$ and $`\widehat{e}_\phi `$ for a generic nanotube, which is encoded in the chiral angle $`\eta `$. The constant phases $`e^{\pm i\eta }`$ can be absorbed by appropriately redefining the operators as $`F_{A2}e^{i\eta }F_{A2}`$ and $`F_{B1}e^{i\eta }F_{B1}`$, and the final result can be written down in the form $$H_{so}=d^2\stackrel{}{r}F^{}\left[u_{}T_0\tau _1\sigma _2+u_{}\mathrm{sin}(y/R)T_3\tau _2\sigma _1\right]F,$$ (21) with $`u_{}=u_{}=3du/2`$. For the sake of generality, we continue to use different coupling constants $`u_{}`$ and $`u_{}`$. It is worthwhile to mention that the leading term for the Rashba spin-orbit coupling in a CNT, Eq. (21), does not depend on longitudinal momentum. This is due to the peculiar band structure of graphene with its isolated Fermi ($`K`$) points. In the above derivation, we also find terms that are linear in momentum, i.e., contain spatial derivatives of the electron operators. Such terms only produce tiny renormalizations of the velocities and will be neglected here. The second term in Eq. (21) allows for spin flips and mixes transverse subbands. From now on, for simplicity, we consider just a single Fermi point, say, $`K`$. After the global $`SU(2)`$ rotation $`\sigma _1\sigma _2\sigma _3`$ in spin space, we get in compact notation $`_0`$ $`=`$ $`v\left[i\tau _1_yi\tau _2_x\right].`$ (22) $`_{so}`$ $`=`$ $`u_{}\tau _1\sigma _3+u_{}\mathrm{sin}(y/R)\tau _2\sigma _2.`$ (23) Note that the exact spectrum of $`_0+_{so}`$ with $`u_{}=0`$ can be obtained straightforwardly. In general, however, due to the smallness of the SO coupling (see below), it is enough to treat $`_{so}`$ perturbatively. The following detailed derivation is then necessary to correctly evaluate the effect of the SO coupling, and moreover it is interesting and important for the generalization to the interacting case, and for the analysis of features involving the electron wavefunction (as for instance electron-phonon interactions). The eigenvalues of $`_0`$ are given by $$ϵ_{an\sigma }(q)=av\sqrt{k_{}^2(n)+q^2}aϵ_n(q),$$ (24) where $`k_{}(n)=(n+n_0)/R`$ denotes the transverse momentum, $`q`$ the longitudinal one, $`a=\pm `$ labels the conduction/valence band, and $`\sigma =\pm `$ the spin. Here $`n_0=0`$ for intrinsically metallic shells, but generally it can be taken as $`0n_01/2`$ to take into account chirality gaps or orbital magnetic fields along $`\widehat{X}`$. The transverse subbands are labeled by integer values $`n=0,1,2,\mathrm{},𝒩1`$, where $`𝒩=2(N^2+M^2+NM)/\mathrm{gcd}(2M+N,2N+M)`$ for ($`N,M`$) tubes . $`𝒩`$ is typically much larger than the actual number $`N=[k_FR]`$ of occupied subbands, where we define $`k_F=\mu /v`$ with the doping level $`\mu `$ that we assume positive here. The velocity $`v_n`$ for electrons in subband $`n`$ at the Fermi level (in the absence of $`H_{so}`$) and the corresponding Fermi momentum $`q_n`$ are then given by $$v_n=v\sqrt{1[(n+n_0)/(k_FR)]^2},q_n=k_Fv_n/v.$$ (25) The eigenvalues (24) are spin-independent and thus doubly degenerate. The corresponding eigenstates are denoted $`|nqa\sigma `$, where $`|n`$ and $`|q`$ are respectively plane waves in circumferential and longitudinal direction. In coordinate representation they read $$\psi _{nqa\sigma }(x,y)x,y|nqa\sigma =\frac{e^{ik_{}(n)y}}{\sqrt{2\pi R}}e^{iqx}\xi _{na}(q)\chi _\sigma ,$$ (26) with the bispinor (in sublattice space) $$\xi _{n,a=\pm }(q)=\frac{1}{\sqrt{2}}\left(\begin{array}{c}e^{i\theta _n(q)/2}\\ \pm e^{i\theta _n(q)/2}\end{array}\right),e^{i\theta _n(q)}=\frac{v(k_{}(n)iq)}{ϵ_n(q)}.$$ (27) A different, and here more convenient basis is given by the sublattice states $`|nqp\sigma `$. Their coordinate representation is $$\psi _{nqp\sigma }(x,y)=\frac{e^{ik_{}(n)y}}{\sqrt{2\pi R}}e^{iqx}\xi _p\chi _\sigma ,$$ (28) where $`p=A,B`$ and $$\xi _A=\left(\begin{array}{c}1\\ 0\end{array}\right),\xi _B=\left(\begin{array}{c}0\\ 1\end{array}\right).$$ Their usefulness stems from the fact that the $`|nqp\sigma `$ can be factorized as $$|nqp\sigma =|n|q|p\sigma ,|p\sigma =\xi _p\chi _\sigma ,$$ (29) where $`|p\sigma `$ is independent of $`n`$ and $`q`$. Using this basis, we can expand the field operator $`F(\stackrel{}{r})`$ on the tube surface as $$F(\stackrel{}{r})=\underset{n,p,\sigma }{}\frac{dq}{2\pi }\psi _{nqp\sigma }(x,y)c_{np\sigma }(q)=\underset{n}{}F_n(x)y|n,$$ (30) where the operator $`c_{np\sigma }(q)`$ destroys an electron in the state $`|nqp\sigma `$, and we introduce the 1D field operators $`F_n(x)`$. Alternatively, using the basis of eigenstates of $`H_0`$, $`F(\stackrel{}{r})`$ can be expanded as $$F(\stackrel{}{r})=\underset{n,a,\sigma }{}\frac{dq}{2\pi }\psi _{nqa\sigma }(x,y)c_{na\sigma }(q),$$ (31) where the operators $`c_{na\sigma }(q)`$ destroy conduction ($`a=+`$) or valence ($`a=`$) electrons with spin $`\sigma `$ in subband $`n`$. Notice that in what follows the spin index is left implicit. The relation between the operators $`c_{na}`$ and $`c_{np}`$ is easily found to be $$\left(\begin{array}{c}c_{n+}(q)\\ c_n(q)\end{array}\right)=\frac{1}{\sqrt{2}}\left(\begin{array}{cc}e^{i\theta _n(q)/2}& e^{i\theta _n(q)/2}\\ e^{i\theta _n(q)/2}& e^{i\theta _n(q)/2}\end{array}\right)\left(\begin{array}{c}c_{nA}(q)\\ c_{nB}(q)\end{array}\right).$$ (32) We now proceed by treating the spin-orbit hamiltonian using perturbation theory. First, we diagonalize $`H_0\mu N`$ for a fixed transverse subband $`n`$, $`H_0^{(n)}\mu N^{(n)}`$ $`=`$ $`v{\displaystyle 𝑑xF_n^{}[k_{}(n)\tau _1+(i_x)\tau _2\mu ]F_n}`$ $`=`$ $`{\displaystyle \underset{a=\pm }{}}{\displaystyle \frac{dq}{2\pi }[aϵ_n(q)\mu ]c_{na}^{}c_{na}}.`$ Next we expand around the Fermi points $`\pm q_n`$ defined in Eq. (25), which introduces right- and left-movers, $`r=\pm =R/L`$, as the relevant low-energy degrees of freedom. For small deviations $`k`$ from $`\pm q_n`$, Taylor expansion yields $`ϵ_n(\pm q_n+k)\mu \pm v_nk`$, where $`v_n`$ is given in Eq. (25). Since we assumed $`\mu >0`$, we may now restrict ourselves to the conduction band, $`a=+`$. For the hamiltonian, we then obtain $`H_0^{(n)}\mu N^{(n)}`$ $`=`$ $`{\displaystyle \underset{r=\pm }{}}v_n{\displaystyle \frac{dk}{2\pi }(rk)c_{nr}^{}(k)c_{nr}(k)}`$ $`=`$ $`{\displaystyle \underset{r=\pm }{}}v_n{\displaystyle 𝑑x\psi _{nr}^{}(ir_x)\psi _{nr}},`$ where $`c_{nr}(k)c_{n+}(rq_n+k)`$ and $`\psi _{nr}(x)=\frac{dk}{2\pi }e^{ikx}c_{nr}(k)`$. This introduces $`R/L`$-moving 1D fermion operators for each subband $`n`$ (and spin $`\sigma `$). The relation of these 1D fermions with the original operator $`F_n(x)`$ is given by $`F_n(x)`$ $`=`$ $`e^{iq_nx}{\displaystyle \frac{dk}{2\pi }\frac{e^{ikx}}{\sqrt{2}}\left(\begin{array}{c}e^{i\theta _n(q_n)/2}\\ e^{i\theta _n(q_n)/2}\end{array}\right)c_{nR}(k)}`$ (35) $`+`$ $`e^{iq_nx}{\displaystyle \frac{dk}{2\pi }\frac{e^{ikx}}{\sqrt{2}}\left(\begin{array}{c}e^{i\theta _n(q_n)/2}\\ e^{i\theta _n(q_n)/2}\end{array}\right)c_{nL}(k)}.`$ (38) Notice that, while in general the unitary transformation from sublattice space to the conduction/valence band description depends on longitudinal momentum, in the continuum limit, one can use the transformation directly at the Fermi momenta. This is consistent with the neglect of band curvature effects implicit in the linearization of the dispersion relation, which is unproblematic away from van Hove singularities associated with the onset of new subbands . At these points, the concept of $`R/L`$-movers breaks down, and some of our conclusions below may change. Next we express the Rashba hamiltonian (23) in terms of $`R/L`$-movers. The first term results in $$H_{so}^{}=\frac{u_{}v}{\mu }\underset{nr}{}k_{}(n)\frac{dk}{2\pi }c_{nr}^{}(k)\sigma _3c_{nr}(k).$$ (39) The presence of the factor $`k_{}(n)`$ results from a careful treatment of the phases in Eq. (35). In Eq. (39) we omit an additional term mixing right- and left-movers. This term contains a rapidly oscillating factor $`e^{\pm 2iq_nx}`$ and therefore is strongly suppressed by momentum conservation. The second term in Eq. (23) again contains the oscillating phase factor $`e^{\pm i(q_n\pm q_{n+1})x}`$, which leads to a drastic suppression of $`H_{so}^{}`$ at low energies and long wavelengths. Of course, this argument relies in an essential way on the smallness of the coupling $`u_{}`$, as one expands around the hamiltonian $`H_0`$. We conclude that away from van Hove singularities, the only important Rashba term is given by $`H_{so}^{}`$ in Eq. (39). This term has the appearance of a static homogeneous but channel-dependent magnetic field. ## 4 Oscillatory TMR effects in nanotubes In this section we will analyze the consequences of our findings regarding spin-orbit couplings in CNTs, see Eqs. (39), for the observability of spin precession effects encoded in the Datta-Das oscillations of the TMR. Based on our expressions, it is possible to estimate the order of magnitude of this effect. For a concrete estimate, let us put $`E_0=0.2eV_G/(\kappa D)`$, where $`D`$ is the gate-tube distance, $`V_G`$ the gate voltage, and $`\kappa `$ denotes the dielectric constant of the substrate. For a given channel $`n`$, the Rashba-induced energy splitting is then easily estimated as $$\frac{\mathrm{\Delta }E_n}{eV_G}=(\gamma _1+\gamma _2)\frac{0.6dv}{\mu }\frac{|n+n_0|}{R}\frac{\alpha ^2\beta d^4\lambda _c^2}{4\kappa D},$$ where $`\lambda _c=\mathrm{}/mc=3.86\times 10^{13}`$ m is the Compton length. Plugging in the definition of $`\alpha ,\beta `$, we get $$\frac{\mathrm{\Delta }E_n}{eV_G}=\frac{0.6(\gamma _1+\gamma _2)}{256\pi \kappa }(d/a_0)^5\frac{\lambda _c^2}{Da_0}\frac{|n+n_0|}{k_FR}.$$ (40) Bands with small $`n`$ are only weakly split, and hence do not contribute to oscillatory TMR behavior. This argument suggests that Datta-Das oscillations in principle could survive in a CNT, even when there are many channels. The major contribution will come just from the few bands with the largest $`n`$. To estimate the accumulated phase difference due to the different precession length of the two split eigenstates, let us put $`(n+n_0)/(k_FR)1`$, which represents the dominant contribution, and set $`\kappa =1`$. Then Eq. (40) gives as order-of-magnitude estimate $$\mathrm{\Delta }E/(eV_G)2\times 10^6a_0/D.$$ (41) Even when assuming a very close-by gate, this gives only a tiny splitting, in retrospect justifying perturbation theory. This splitting now translates into a momentum splitting $`\mathrm{\Delta }k_n=\mathrm{\Delta }E/v_n`$, and hence into a precession phase mismatch along the CNT of length $`L`$ . For the $`n`$th band, this phase difference is $$\mathrm{\Delta }\varphi _n=\mathrm{\Delta }k_nL2\times 10^6\frac{L}{D}\frac{eV_G}{\mathrm{}v_n/a_0}.$$ (42) This phase difference should be of order $`2\pi `$ to allow for the observation of Datta-Das oscillatory TMR effects . Away from a van Hove singularity, Eq. (42) predicts that oscillations appear on a gate voltage scale of the order of $`10^6`$ to $`10^7`$ V for $`LD`$, which would make Datta-Das oscillations unobservable. This argument also shows that this interpretation can be ruled out for the parameters relevant for the Basel experiment . From Eq. (42), we can then suggest several ways to improve the situation. First, one should use very long CNTs, while at the same time keeping the gate very close, and second, an enhancement can be expected close to van Hove singularities. Of course, very close to a van Hove singularity, some of our arguments above break down, but the general tendency can nevertheless be read off from Eq. (42). Furthermore, electron-electron interactions can also enhance spin-orbit effects . To conclude, we have presented a detailed microscopic derivation of Rashba spin-orbit coupling in carbon nanotubes. It turns out that the Rashba SO coupling is small, and therefore the prospects for observing spin-precession effects like Datta-Das oscillations in the tunneling magnetoresistance are not too favorable. However, for very long CNTs, close-by gates, and in the vicinity of a van Hove singularity, the requirements for observability of these effects could be met in practice. We thank T. Kontos and C. Schönenberger for motivating this study. Support by the DFG (Gerhard-Hess program) and by the EU (DIENOW network) is acknowledged. ## References
warning/0506/hep-ex0506024.html
ar5iv
text
# TOP PHYSICS AT THE LHC ## 1 Introduction The Large Hadron Collider (LHC) is expected to start operation in 2007. It will accelerate proton beams and bring them to collision at a centre of mass energy of $`\sqrt{s}`$=14 TeV and at luminosities between $`(12)\times 10^{33}cm^2s^1`$ (initial ’low-luminosity’ phase) and $`10^{34}cm^2s^1`$ (’high-luminosity’). Top quark physics studies will be accomplished by the general purpose experiments ATLAS and CMS. The reasons to study the physics of top quarks are numerous. The top quark is the heaviest known fundamental particle – its mass being a fundamental parameter of the Standard Model – and the only quark that does not hadronize because of its short lifetime ($`\mathrm{\Gamma }_t>>\mathrm{\Lambda }_{QCD}`$). The top quark mass $`m_t`$ is of particular interest, because it is related through radiative corrections (the so-called $`\mathrm{\Delta }r`$ parameter) to the mass of the W and Higgs bosons, $`m_W`$ and $`m_H`$. Precise measurements of $`m_t`$ and $`m_W`$ thus allow to constrain $`m_H`$. Top quark events will also be a major source of background for many search channels. They can be also very useful to calibrate the detector and reconstruction algorithms, e.g. the jet energy scales and b-tagging performance. At the LHC, gluon-gluon fusion is the dominant process for $`t\overline{t}`$ production, contributing about 87% of the production cross section, whereas quark-antiquark annihilation contributes only about 13%. This is contrary to the situation at the TEVATRON. The $`t\overline{t}`$ production cross-section is about $`\sigma _{t\overline{t}}800pb`$. This implies that during the low luminosity phase of the LHC ($`10^{33}cm^2s^1`$), about one $`t\overline{t}`$ pair will be produced per second, resulting in about $`10^7`$ $`t\overline{t}`$ pairs per year. Because of $`|V_{tb}|1`$, the top quark decays almost exclusively into a W boson and a b quark, $`BR(tWb)1`$. The top quark decay is thus characterised through the decay of the W boson. In $`t\overline{t}`$ events, about 5% fall into the di-lepton category, $``$ 30% into the semi-leptonic one and $``$ 45% into the fully hadronic one (only electrons and muons are considered as leptons here). ## 2 Measurements of the top quark mass The semi-leptonic channel $`t\overline{t}b\overline{b}l\nu q\overline{q}`$ is the most promising for an accurate measurement of the top quark mass. It has a relative large branching ratio and the isolated lepton ($`e`$ or $`\mu `$) allows efficient triggering. A typical analysis requires an isolated lepton, missing transverse energy, and at least four jets within the tracker acceptance. To suppress backgrounds and to resolve the combinatorics within the event, one or two jets have to be tagged as b-jets. The mass of the hadronically decaying top quark, thus the mass of the $`Wb=jjb`$ system, is used for the measurement. Figure 1 shows the reconstructed top masses for ATLAS $`^\mathrm{?}`$ and CMS $`^\mathrm{?}`$. The remaining backgrounds (mainly W and Z boson(s) + jets) are very low. Already after one year of running at low luminosity, the measurement uncertainty is clearly dominated by systematics. Of particular importance is the knowledge of the b-jet energy scale (the sensitivity to the light jet scale is strongly reduced by using the W boson mass as constraint), which has to be determined at the level of a percent to reach the desired accuracy. It seems to be feasible to reach a precision on the top mass of about 2 GeV, maybe an ultimate precision of about 1 GeV could be reached when all effects are well understood. The large number of events available allows to measure the top mass in different sub-samples using different methods showing different sources of systematic errors. ATLAS $`^\mathrm{?}`$ has studied semi-leptonic events, where the top quarks are produced almost back-to-back with large transverse momentum, leading to very collimated top decay products well separated in hemispheres. The top quark mass is computed from the calorimeter towers within a large cone. The dominant systematics using this method is then the contribution from the underlying event and pile-up events. The top mass can also be measured in the fully leptonic and fully hadronic channels. The Di-Lepton channel ($`e`$, $`\mu `$) is very clean. However, the branching ratio is quite small and because of the presence of two neutrinos in the final state the kinematics is underconstrained. The fully hadronic channel is experimentally very challenging. Efficient triggering is difficult and the QCD background is enormous. However, this channel has the biggest branching ratio ($`45\%`$) and the kinematics is fully constrained. Accuracies of about 2-3 GeV seem feasible for these channels $`^\mathrm{?}`$. ATLAS $`^\mathrm{?}`$ and CMS $`^\mathrm{?}`$ also studied the possibility to measure the top mass in the semi-leptonic channel in final states with an exclusively reconstructed $`J/\mathrm{\Psi }`$ from a b-hadron decay in a b-jet, where the $`J/\mathrm{\Psi }`$ carries a large fraction of the b-hadron momentum because of its large mass. The mass of the $`J/\mathrm{\Psi }l`$ system is strongly correlated to the top mass. The method is thus rather insensitive to the jet energy scale, the understanding of the b-fragmentation becoming crucial. Because of the tiny branching ratio of this final state, this analysis has to be carried out during the high luminosity phase of the LHC. Even after some 100$`fb^1`$, the statistical error is expected to contribute significantly to the total error ($``$ 1 GeV for a total error of $``$ 1.5 GeV). ## 3 $`t\overline{t}`$ production properties A measurement of the $`t\overline{t}`$ production cross section allows to test QCD calculations. Differential cross sections (e.g. $`d\sigma _{t\overline{t}}/d\eta `$) give access to parton distribution functions (PDFs). Heavy particles decaying into a $`t\overline{t}`$ pair could show up as resonances in the $`t\overline{t}`$ mass spectrum $`d\sigma _{t\overline{t}}/dm_{t\overline{t}}`$ and thus be an indication for new physics. Furthermore, the $`t\overline{t}`$ production cross section is sensitive to the top quark mass, $`\sigma _{t\overline{t}}1/m_t^2`$. Because of its short lifetime, the top quark does not form bound states and thus there should be no dilution of the spin information in the decay. Top quarks are not expected to be produced polarised but their spins are expected to be correlated. The asymmetry $`𝒜=\frac{N(t_L\overline{t}_L+t_R\overline{t}_R)N(t_L\overline{t}_R+t_R\overline{t}_L)}{N(t_L\overline{t}_L+t_R\overline{t}_R)+N(t_L\overline{t}_R+t_R\overline{t}_L)}`$ allows to distinguish between the g-g fusion and q-$`\overline{\text{q}}`$ production processes ($`𝒜0.431`$ vs. $`𝒜0.469`$). In fully leptonic events, $`𝒜`$ can be extracted from a fit to the double differential distribution of the angles between the leptons in their respective top frame and the top quark in the $`t\overline{t}`$ rest frame. A CMS study $`^\mathrm{?}`$ shows that the following accuracies can be reached after 30$`fb^1`$ of integrated luminosity: $`\sigma (𝒜)_{stat.}0.035`$, $`\sigma (𝒜)_{syst.}0.028`$. ## 4 Top quark properties The large number of $`t\overline{t}`$ events will allow precise measurements of top quark properties and tests of the V-A structure of charged current weak interactions. The top quark charge has not yet been experimentally confirmed to be $`2/3`$. Two approaches have been studied by ATLAS $`^\mathrm{?}`$ to distinguish between $`t(Q=2/3)W^+b`$ and $`t(Q=4/3)W^{}b`$. The first approach investigates $`t\overline{t}\gamma `$ events. Since the radiation of photons in these events is roughly proportional to $`Q_t^2`$ for the production process (radiation can also occur in the decay), the $`p_t`$ spectrum of the photons in $`t\overline{t}\gamma `$ allows to distinguish the two top quark charge values. The second approach measures the charges of all top quark decay products. Whereas for the leptonic W boson decay the charge is given by the charge of the lepton ($`e`$ or $`\mu `$), a jet-by-jet estimation of the b-quark charge is not possible. However, quantities correlated to the charge of the b-quark allow to distinguish between the cases on a statistical basis. As an example, the jet charge $`Q_{jet}=\frac{_iq_i|\stackrel{}{jet}\stackrel{}{p}_i^\kappa |}{_i|\stackrel{}{jet}\stackrel{}{p}_i^\kappa |}`$ has been studied. Both methods should allow an unambiguous measurement of the top quark charge after one year of data taking at low luminosity (10 $`fb^1`$). Because of the V-A structure of weak charged current interactions, the W bosons in top quark decays are expected to be polarised. The predicted population of the helicity states of the W boson are: $`h_W(1)0.3`$, $`h_W(0)0.7`$, $`h_W(+1)0`$. In semi-leptonic events, the angle between the lepton in the W boson rest frame and the W boson in the top quark frame, $`\mathrm{\Theta }_l^{}`$, can be analysed. A CMS study $`^\mathrm{?}`$ gives the following accuracies for the longitudinal helicity state after an integrated luminosity of 10 $`fb^1`$: $`\sigma (h_W=0)_{stat.}0.023`$, $`\sigma (h_W=0)_{syst.}0.022`$. ## 5 Single top production Single top quark events can be created via electro-weak processes. The contributing diagrams together with their expected cross-sections are shown in Figure 2. Single top events give direct access to the absolute value of the CKM matrix element $`V_{tb}`$. Backgrounds ($`t\overline{t}`$, Wjj, Wbb) are more severe than for $`t\overline{t}`$ events and the ability to extract clear signals depends more critically on the detector performance. The different production processes show different final states and are thus typically studied separately. To reject the enormous QCD background, the leptonic decay channel $`tbl\nu `$ ($`l=e,\mu `$) is studied. An ATLAS study $`^\mathrm{?}`$ gives the following statistical experimental errors on $`|V_{tb}|`$ after $`30fb^1`$: $`\sigma (|V_{tb}|)_{(Wg)}0.4\%`$, $`\sigma (|V_{tb}|)_{(Wt)}1.4\%`$, $`\sigma (|V_{tb}|)_{(W^{})}2.7\%`$. The dominating systematic uncertainties are expected to come from theoretical uncertainties of the cross section calculations and the luminosity measurement. ## 6 Calibration with $`t\overline{t}`$ events The possibility to isolate clean samples of $`t\overline{t}`$ events allows the calibration of reconstruction algorithms. Using the W boson mass as a constraint allows to calibrate the jet energy scale of light quark jets. Jet samples enriched in b-quark jets (from the top quark decay) and light quark jets can be isolated and used to determine the b-tagging efficiencies on the data themselves. ## 7 $`t\overline{t}`$ in associated Higgs production For low Higgs masses ($`m_H135`$ GeV), the dominant decay mode of the Higgs boson is $`Hb\overline{b}`$. However, this decay mode can only be exploited in channels of associated Higgs production, $`ttH`$ being the most promising one. From the experimental point of view, this channel is complementary to the channel $`H\gamma \gamma `$. Furthermore, this channel would allow a measurement of the top-Higgs Yukawa coupling. In order to find the b quark jets from the Higgs boson decay, the $`t\overline{t}`$ system can be fully reconstructed. ATLAS $`^\mathrm{?}`$ and CMS $`^\mathrm{?}`$ studies indicate that this channel might support a discovery of a light Higgs boson. ## 8 Conclusions At the LHC, top quarks will be produced abundantly. The goal for the top mass measurement is to reach an accuracy of approximately 1 GeV. The large available statistics will also allow to determine many top quark properties both in production and decay. Single top production will make a precise determination of $`|V_{tb}|`$ possible. ## References
warning/0506/astro-ph0506526.html
ar5iv
text
# The Origin of T Tauri X-ray Emission: New Insights from the 𝐶⁢ℎ⁢𝑎⁢𝑛⁢𝑑⁢𝑟⁢𝑎 Orion Ultradeep Project ## 1 Introduction ### 1.1 X-ray emission from young stellar objects Young stellar objects (YSOs) in all evolutionary stages from class I protostars to ZAMS stars show highly elevated levels of X-ray activity (for recent reviews on the X-ray properties of YSOs and on stellar coronal astronomy in general see Feigelson & Montmerle 1999 and Favata & Micela 2003). X-ray observations of star forming regions allow to study the high energy processes in YSOs, which are of great importance for our understanding of the star formation process. For example, the X-ray emission from a YSO should photoionize its circumstellar material and thus influence accretion as well as outflow processes, both of which are thought to be based on the interaction of ionized material with magnetic fields. The X-ray emission from the central YSO is certainly an important, probably even the dominant factor in determining the ionization structure of protoplanetary disks, and has therefore a strong impact on processes like the formation of proto-planets (e.g. Glassgold, Feigelson & Montmerle, 2000; Matsumura & Pudritz, 2003). The first discoveries of X-ray emission from T Tauri stars (TTS, $`=`$ low-mass pre-main sequence (PMS) stars) were made with the EINSTEIN X-ray observatory (e.g. Feigelson & DeCampli, 1981) and revealed a surprisingly strong X-ray activity, exceeding the solar levels by several orders of magnitude. The X-ray observations also revealed a new population of young stellar objects (Walter et al., 1988), the “weak-line T Tauri stars” (WTTS), which lack the classical optical signposts of youth, like strong H$`\alpha `$ emission, of the previously known “classical T Tauri stars” (CTTS). The ROSAT observatory increased the number of observed star forming regions, and thereby the number of known X-ray emitting TTS, considerably (e.g. Feigelson et al., 1993; Casanova et al., 1995; Gagné et al., 1995; Neuhäuser et al., 1995; Preibisch, Zinnecker, & Herbig, 1996). The ROSAT All Sky Survey lead to the X-ray detection of extended populations of TTS in and around many star forming regions (e.g. Neuhäuser, 1997) and demonstrated that the stellar populations of star forming regions are considerably larger than suspected by earlier surveys based on classical youth indicators such as H$`\alpha `$ emission. The ROSAT All Sky Survey was also well suited to study the X-ray properties in complete, volume limited samples of nearby field stars. An important result from such studies was that apparently all cool dwarf stars are surrounded by X-ray emitting coronae, with a minimum X-ray surface flux around $`10^4`$ erg/sec/cm<sup>2</sup> Schmitt (1997). The ASCA satellite detected X-ray emission from numerous deeply embedded YSOs; due to its rather poor spatial resolution, however, the proper identification of the X-ray sources was often difficult. While the X-ray missions of the last two decades provided important information about the X-ray properties of YSOs, there were also serious limitations. First, the typical samples of X-ray detected objects in young clusters and star forming regions contained hardly more than $`100`$ objects, too few to allow well founded statistical conclusions to be drawn. Second, a large fraction of the known cluster members (especially low-mass stars) remained undetected in X-rays, and any correlation studies had therefore to deal with large numbers of upper limits. Third, especially in dense clusters, the individual sources could often not be spatially resolved, and so the proper identification of the X-ray sources was difficult or impossible. Finally, only a relatively small number of individual young stars were bright enough in X-rays to allow their spectral and temporal X-ray properties to be studied in detail, and it is not clear whether these stars really are “typical” cases or perhaps peculiar objects. With the advent of $`Chandra`$ and XMM-Newton, the situation has improved substantially. Due to the large collecting areas of these observatories, their sensitivity is at least an order of magnitude better than that of earlier missions. Due to their wide energy band, extending from $`0.20.5`$ keV up to $`810`$ keV, they are very well suited to study the hard X-ray emission from highly obscured YSOs (e.g., Skinner et al., 2003). Furthermore, $`Chandra`$ has a superb point spread function, providing a spatial resolution of better than $`1^{\prime \prime }`$; this abolishes the usual identification problems in nearby star forming regions. ### 1.2 Open questions about the X-ray activity from TTS We enunciate two basic, still unresolved questions concerning the origin of the elevated X-ray activity of TTS: Does the strong X-ray activity of TTS, with X-ray luminosities up to $`10^4`$ times and plasma temperatures up to $`50`$ times higher than seen in our Sun, originate from solar-like coronae? If so, are these coronae created and heated by solar-like (although strongly enhanced) magnetic dynamo processes, or are fundamentally different magnetic structures and heating mechanisms involved? One main obstacle on the way towards an understanding of TTS X-ray activity is the fact that the solar corona has an extremely complex and dynamic structure with many different facets (e.g., Aschwanden et al., 2001); it is not clear to what degree comparisons and extrapolations from the solar to the stellar case make sense. A second problem is that even for the Sun, which can be studied in great detail at high spatial, temporal, and spectral resolution, the important question about the heating of the solar corona remains puzzling, even after 5 decades of intense research (e.g., Walsh & Ireland, 2003). The third problem is our lack of a sound understanding of the dynamo processes which are the ultimate origin of the magnetic activity in the Sun and in stars (e.g., Ossendrijver, 2003). The X-ray activity of main-sequence (MS) stars is mainly determined by their rotation rate. The well established rotation–activity relation (e.g. Pallavicini et al., 1981; Pizzolato et al., 2003) is given by the power-law relation $`L_\mathrm{X}/L_{\mathrm{bol}}P_{\mathrm{rot}}^{2.6}`$, in agreement with the expectations from solar-like $`\alpha \mathrm{\Omega }`$ dynamo models (e.g. Maggio et al., 1987). At periods shorter than $``$ 2-3 days, the activity saturates at $`\mathrm{log}\left(L_\mathrm{X}/L_{\mathrm{bol}}\right)3`$ for reasons that are not yet understood. The plasma temperatures generally increase with the level of X-ray activity, scaling roughly as $`T_\mathrm{X}\left(L_\mathrm{X}/L_{\mathrm{bol}}\right)^{0.5}`$ (e.g. Preibisch, 1997). Most TTS rotate quite rapidly, and neither their X-ray luminosities nor their plasma temperatures are unusual when compared to rapidly rotating MS stars. However, a relation between rotation and X-ray activity could never be convincingly established for TTS; in most studies the small number of X-ray detected TTS with known rotation periods did not allow to draw sound conclusions. This problem of small and often biased samples has, however, recently been overcome with two $`Chandra`$ studies of the ONC (Feigelson et al., 2002a; Flaccomio et al., 2003b), both of which found no significant relation between X-ray activity and rotation. This strongly puts into question the solar-like dynamo activity scenario for TTS. Another argument against solar-like dynamos comes from theoretical considerations: at ages of only a few Myr, the TTS are usually thought to be fully convective, and therefore the standard solar-like $`\alpha \mathrm{\Omega }`$ dynamo, which is anchored at the boundary between the convective envelope and the inner radiative core, should not work. Theoreticians have developed alternative dynamo concepts (e.g., Küker & Rüdiger 1999; Giampapa et al. 1996) that may work in fully convective stars. A problem with these and other models is that they usually do not make quantitative predictions that can be easily tested from observations. Further possibilities for the origin of X-ray emission from TTS include magnetic fields coupling the stars to their surrounding circumstellar disk (see, e.g., Hayashi et al. 1996; Montmerle et al. 2000; Isobe et al., 2003; Romanova et al., 2004), or X-ray emission from accretion shocks (see, e.g., Kastner et al., 2002; Stelzer & Schmitt, 2004; Favata et al., 2003, 2005). The investigation of these possibilities in the light of the the Chandra Orion Ultradeep Project (COUP) (catalog ADS/Sa.CXO#obs/COUP) data will be a major topic of this study. ### 1.3 Properties of the ONC and previous X-ray observations The Orion Nebula is an HII region on the near side of a giant molecular cloud, which contains one of the most prominent and nearby ($`D450`$ pc) star forming regions (for a recent review see O’Dell, 2001). This star forming region contains a massive cluster of young ($`10^6`$ yr) stars (cf. Herbig & Terndrup, 1986; McCaughrean & Stauffer, 1994; Hillenbrand, 1997), which is known as the Orion Nebula Cluster (ONC). The Orion Nebula is illuminated mainly by the two O-type stars $`\theta ^1\mathrm{Ori}`$ C and $`\theta ^2\mathrm{Ori}`$ A. Since the ONC is a perfect laboratory for observations of star formation over the full stellar mass range, it is one of the best investigated star forming regions and has been observed at virtually any wavelength. Hillenbrand (1997) has compiled a catalog of nearly 1600 optically visible stars within $`2.5`$ pc of the Trapezium; for over 900 of these stars enough information is avaliable to place them into the HR-diagram and to determine their masses and ages by comparison with theoretical PMS evolution models. The ONC has been observed with basically all previous X-ray observatories (see e.g. Ku & Chanan (1979) for EINSTEIN observations; Gagné et al. (1995); Geier et al. (1995); Alcalá et al. (1996) for ROSAT observations; Yamauchi et al. (1996) for ASCA observations). However, the high spatial density of stars in the ONC and the poor spatial resolution of these X-ray observatories did not allow a reliable identification of many X-ray sources. Only $`Chandra`$ with its superb point spread function is suitable for studying the ONC, where the mean separation between the sources in the inner $`1^{}`$ radius area is only $`5^{\prime \prime }`$. The ONC has been observed with both imaging instruments onboard of $`Chandra`$. The results of two ACIS-I observations with a combined exposure time of 23 hours were reported in Garmire et al. (2000) and Feigelson et al. (2002a, b, 2003). 1075 individual sources were detected, 91% of which could be identified with known stellar members of the cluster. Flaccomio et al. (2003a, b) presented the analysis of a 17.5 hr HRC-I observation of the ONC, which yielded 742 X-ray sources in the $`30^{}\times 30^{}`$ field-of-view. Furthermore, some of the brightest X-ray sources in the ONC have also been studied with the High Energy Transmission Grating Spectrometer (Schulz et al., 2000, 2001, 2003), but most of these sources are massive stars which are not the topic of this paper. In this paper, we discuss the X-ray data on the TTS in the ONC resulting from COUP, by far the longest and most sensitive X-ray observation ever obtained for the ONC. The plan of this paper is as follows: after briefly describing the COUP observation in §2, we define in §3 the optical sample which will be the basis of our studies, and then investigate the relation of the X-ray emission to basic stellar parameters in §4. In §5 we study in detail the relation between X-ray emission, rotation and convection. In §6 we discuss the origin of the large scatter seen in the correlations between X-ray activity and other stellar parameters. In §7, we investigate the plasma temperatures as determined from the fits to the X-ray spectra. Section 8 deals with the possible connections between X-ray emission and circumstellar accretion disks. Finally, in §9 we discuss the implications of our results with respect to the origin of the TTS X-ray emission. ## 2 The COUP observation The COUP observation is the deepest and longest X-ray observation ever made of a young stellar cluster, providing a rich and unique dataset for a wide range of science studies. The observational details and a complete description of the data analysis can be found in Getman et al. (2005a), here we summarize only the aspects that are most important to our studies. The COUP observation was performed between 8 Jan 2003 and 21 Jan 2003, utilizing the Advanced CCD Imaging Spectrometer (ACIS) in its imaging configuration, which gives a field of view of $`17^{}\times 17^{}`$. The total exposure time of the COUP image was 838 100 sec (232.8 hours or 9.7 days). The spatial resolution of ACIS is better than $`1^{\prime \prime }`$ over most of the field of view. The very low background allows the reliable detection of sources with as little as $`5`$ source counts. The final COUP source catalog lists 1616 individual sources. The superb point spread function and the high accuracy of the aspect solution allowed a clear and unambiguous identification of nearly all X-ray sources with optical or near-infrared counterparts<sup>1</sup><sup>1</sup>1The median offsets between COUP sources and near-infrared or optical counterparts are only 0.15″and 0.24″, respectively. Spectral analysis was performed using a semi-automated approach to produce an acceptable spectral model for as many as possible sources. The XSPEC spectral fitting programme was used to fit the extracted spectra with one- or two-temperature optically thin thermal plasma MEKAL models assuming 0.3 times solar abundances and X-ray absorption. The parameters derived in these fits are the hydrogen column density $`N_\mathrm{H}`$ as a measure of the X-ray absorption, and the temperatures $`T_\mathrm{X}`$ and emission measures $`EM`$ of the one or two spectral components. The spectral fitting results were also used to compute the intrinsic (extinction-corrected) X-ray luminosity by integrating the model source flux over the $`0.58`$ keV band. Our analysis in this paper is based on the tabulated X-ray properties and identifications of the COUP sources as listed in Getman et al. (2005a). We use the identifications of the X-ray sources with optical counterparts as given in their Table 9, the X-ray luminosities and X-ray spectral properties as listed in their Tables 8 and 6, and upper limits of undetected stars in the Hillenbrand sample as given in Table 11. The temporal behavior of COUP sources is often very complex, with high-amplitude, rapidly changing flares superposed on apparent quiescent or slowly variable emission as studied in detail by Wolk et al. (2005), Favata et al. (2005) and Flaccomio et al. (2005). For the purpose of the present study, the individual features of the lightcurves are not of interest. We note that the X-ray properties tabulated in the COUP tables, i.e. the count rates, derived spectral parameters, and X-ray luminosities, represent the average over the 10 days exposure time of our dataset. This implies that the effect of short excursions in the X-ray lightcurves, like flares with typical timescales of a few hours, are strongly “smoothed out”. Since most of the young stars are rather fast rotators with periods of less than 10 days, the tabulated X-ray properties represent for these objects the average over at least one rotation period and therefore also smooth out possible rotational modulation. Nevertheless, it would be interesting to establish the level of “quiescent” X-ray emission in the sources, i.e. the sustained, or “typical” level of X-ray emission outside the periods of flares or otherwise elevated activity. For this purpose, we used the results of the Maximum Likelihood Blocks (MLB) lightcurve analysis (Flaccomio et al., 2005), which segments the lightcurves into contiguous sequences of constant count rates and allows to discern between periods of flaring and more constant, sustained X-ray emission. A strict and fully convincing definition for “quiescent” X-ray emission is not possible, especially since any apparently quiescent emission may, in reality, just be a superposition of numerous unresolved small flares. Wolk et al. (2005) empirically establish a proxy for the quiescent emission levels by determining a ”characteristic level” in each lightcurve, which is essentially defined as the average count rate over periods where the count rate is not significantly elevated. An estimate for the characteristic X-ray luminosity can then be obtained by multiplying the temporally averaged X-ray luminosity, as determined from the spectral analysis, with the ratio of the characteristic countrate from the MLB analysis to the mean countrate over the COUP exposure. We note that this simple scaling procedure is not fully self-consistent, because it does not take into account that the X-ray spectral parameters (and thereby the transformation factor from count rate to luminosity) can change as a function of the emission level, but it should be appropriate for our purposes. The difference between the average and characteristic X-ray luminosities is generally not large: the median value of the correction factor is 0.78; only for 13% of the TTS this factor is $`<1/2`$, and for only 4% of the TTS $`<1/3`$. We will show below that the choice of either the average or the characteristic X-ray luminosities has generally very little effect on the observed relations. We will therefore mainly use the temporally averaged X-ray luminosities and consider the characteristics luminosities only in a few cases. ## 3 Definition of the optical sample and X-ray detection completeness ### 3.1 The optical sample of ONC stars The aim of our study is to investigate the X-ray properties of a homogenous and well defined sample of comprehensively characterized TTS (= young late type \[F–M\] stars). We will therefore not consider the COUP detected brown dwarfs (see Preibisch et al., 2005) or embedded objects (see Grosso et al., 2005), or OBA stars (see Stelzer et al., 2005), although we will sometimes compare the more massive stars with TTS. The basis for the construction of our “optical sample” is the Hillenbrand (1997) \[H97 hereafter\] sample of $`1576`$ optically visible ($`I17.5`$) stars within $`2.5`$ pc ($`20^{}`$) of the Trapezium, for $`934`$ of which optical spectral types are known. We used an updated version of the H97 tables in which for many objects spectral types and other stellar parameters have been revised<sup>2</sup><sup>2</sup>2http://www.astro.caltech.edu/~lah/papers/orion\_main.table1.working. Some of the stars in this area are unrelated field stars lying either in the foreground or the background of the Orion Nebula; these should of course be excluded from our studies. We therefore used the membership information from proper motion studies listed in the table of H97: we consider all stars with membership probabilities $`50\%`$ to be bona-fide members of the ONC, while stars with membership probabilities $`<50\%`$ are considered here to be non-members and excluded from our analysis. Stars with no membership information are considered here as likely cluster members, because contamination by foreground field stars is very small<sup>3</sup><sup>3</sup>3H97 estimate that $`97\%`$ of the $`I17`$ mag stars within about 1 pc of the Trapezium (i.e., roughly the field of view of our COUP image) are ONC members, and contamination by background stars is unlikely due to the large visual extinction in the molecular could immediately behind the ONC. H97 assumes that the optical database is representative of all stars in the ONC region and that the completeness of $`60\%`$ is uniform with radius. 1056 of the H97 stars are within the field-of-view of the COUP observation, 892 of which are detected as X-ray sources. Excluding the 33 stars that are identified as non-members, we have 1023 likely ONC members from the H97 sample in the COUP field-of-view and detect 870 of these (i.e., 85%) as X-ray sources. For the analysis in this paper we use those of these stars for which spectral types are known. Our “optical sample” then consists of 639 optically visible likely ONC members from H97 with known spectral type; 598 of these stars (i.e., 94%) are detected as X-ray sources, 41 remain undetected in the COUP image. The spectral types range from O7 for $`\theta ^1C`$ Ori, the most massive and luminous star in the ONC, down to $``$ M6.5 for objects close to the stellar-substellar boundary (at $`0.075M_{}`$) at the age of the ONC. For 575 stars in the COUP optical sample bolometric luminosities are known, allowing them to be placed into the HR-diagram. Masses and ages were estimated for 536 stars by comparison of their location in the HR-diagram to the theoretical PMS evolutionary tracks from Siess, Dufour, & Forestini (2000). The masses in the COUP optical sample range from $`0.1M_{}`$ (the lowest mass in the Siess, Dufour, & Forestini (2000) models) to $`38M_{}`$ for $`\theta ^1`$ C Ori. The visual extinction is known for 631 of the 639 stars in our optical sample and varies from 0 to $`A_V=11`$ mag, with a mean value of $`A_V=1.55`$ mag. Since the optical sample is magnitude limited, extinction introduces a bias, since intrinsically brighter stars can suffer from more extinction and still be included in the sample than the intrinsically fainter stars. We therefore used an extinction limit to construct a more homogeneous sample: we define as the “lightly absorbed optical sample” those stars for which the optical extinction is known and is $`A_V5`$ mag. The lightly absorbed optical sample consists of 586 stars, 554 of which are detected as X-ray sources in the COUP image. This extinction limit also yields a rather uniform sample with respect to the X-ray detection limit: PIMMS simulations for Raymond-Smith spectra with $`kT=2`$ keV show that the detection limit (i.e. the X-ray luminosity that corresponds to a given number of detected source counts) increases by $`0.39`$ dex when going from zero extinction to $`N_\mathrm{H}=8\times 10^{21}\mathrm{cm}^2`$ ($`A_V=5`$ mag). Since the uncertainty of the X-ray luminosity determinations is similar to this factor, our extinction limited sample does not suffer from a significant extinction-dependent X-ray detection bias. To summarize, our “lightly absorbed optical sample” of 586 stars is not 100% complete (because spectral types are not available for all stars in the ONC), but nevertheless should be a statistically representative sample of the ONC young stellar population with low extinction. The only potential systematic selection effect might be that older ($`10`$ Myr) very-low mass ($`M0.2M_{}`$) stars may be missing; it is, however, unclear whether such an older population of ONC members does exist at all. For the 42 stars in the optical sample which were not detected as X-ray sources in the COUP data, we estimated upper limits to their X-ray luminosities from the tabulated upper limits to their count rates following the procedure outlined in Getman et al. (2005a). ### 3.2 X-ray detection completeness Table 1 lists the COUP X-ray detection fractions for the different spectral types. It is important to note here that most of the non-detections of ONC stars are due to X-ray source confusion in the COUP data; the typical case are close ($`1^{\prime \prime }2^{\prime \prime }`$ separation) binary systems, in which only one of the components is clearly detected as an X-ray source (Getman et al., 2005a). In these cases the object would perhaps have been detected if located at a different position. Since the occurrence of source confusion should not depend on stellar parameters, these objects can be considered as “unobserved” and will be ignored in our analysis. We can thus compute an “effective”, confusion-free detection fraction by removing these non-detections with source confusion from the sample. Then, the effective detection fractions range between 97% and 100% for all spectral classes except the A- and B-type stars. This means that less than 3% of the TTS in the optical sample are undetected because their X-ray emission is below our detection limit. With so few undetected objects, we can look at these stars in detail. The undetected B8 star H97-1892 and the A1 star H97-531 are intermediate mass stars and will be discussed in Stelzer et al. (2005). The three undetected K-type stars not suffering from source confusion are H97-62, H97-489, and H97-9320. H97-62 has a rather large extinction of $`A_V=5.26`$ mag that may have absorbed too much of its X-ray emission. H97-489 has no bolometric luminosity and optical extinction listed in H97. The third object is the K6-star H97-9320, which lies far (2.7 mag) below the ZAMS in the HR-diagram, putting considerable doubt on its membership to the ONC; since the star has also no proper-motion membership information in H97, we suspect it to be a non-member and exclude it from our optical sample. We therefore conclude that all K-type stars in the lightly absorbed optical sample are detected. For the 13 undetected M-type stars not suffering from source confusion we note that two objects have optical extinction exceeding $`A_V=5`$ mag, and 4 objects have no proper-motion membership information in H97 and may therefore perhaps be non-members. Only 8 of the M-type stars among the known proper motion members in the lightly absorbed optical sample remain undetected. The upper limits to the fractional X-ray luminosities of the undetected M-type stars range from $`\mathrm{log}\left(L_\mathrm{X}/L_{\mathrm{bol}}\right)<5.46`$ (for H97-305) to $`\mathrm{log}\left(L_\mathrm{X}/L_{\mathrm{bol}}\right)<4.37`$ (for H97-853). This is considerably below the mean fractional X-ray luminosities of the detected M-type stars of $`\mathrm{log}\left(L_\mathrm{X}/L_{\mathrm{bol}}\right)=3.62`$, but still within the range of fractional X-ray luminosities found for the detected M-type stars, three of which have values below the lowest upper limit for the non-detections. We conclude that the very few undetected M-type stars show low, but not necessarily unusually low levels of X-ray activity. To summarize, we find that COUP detects every optically visible star in the ONC sample except a few of the intermediate mass stars (which are not expected to be intrinsic X-ray emitters) and a few of the M-type stars (some of which may perhaps be non-members). We find no indications for the existence of an “X-ray quiet” population of stars with suppressed magnetic activity. Our analysis is thus based on a (nearly) complete sample, and we can be very confident that our conclusions will not be affected by non-detections. The only remaining concern is about the completeness of the optical sample, which may not be very well established for older ($`10^7`$ yr) very low-mass ($`0.2M_{}`$) stars. It is interesting to note that most (1047 of 1616, i.e. 65%) of the COUP X-ray sources were already detected in the previous 23 hr exposure ACIS observation or the 17.5 hr HRC observation of the ONC. Since the field-of-view covered by the different observations is not identical, we focus on the inner $`8^{}`$ radius area, which is included in all $`Chandra`$ observations considered here. In this area, 970 (66.6%) of the 1457 COUP sources, 475 (90.9%) of the 522 COUP sources in the optical sample, and 438 (91.3%) of the 480 COUP sources in the lightly absorbed optical sample were already detected in the previous 23 hr ACIS observation or the 17.5 hr HRC observation. With more than 10 times the earlier exposure times, COUP leads to the detection of only a relatively small number of 47 (42) new X-ray sources that could be identified with stars in the (lightly absorbed) optical sample. The COUP data nevertheless represent an important step forward over the previous observations, since they increased the X-ray detection fraction in the lightly absorbed optical sample from $`88\%`$ to at least 97%, transforming an incomplete sample to a (nearly) complete sample. Also, the COUP data for the individual sources have much higher S/N (counts per source) than previous data sets, thus allowing a much more reliable determination of the X-ray source properties. ## 4 Relation of the X-ray emission to basic stellar parameters Since the origin of the X-ray activity in TTS is still not well known, it is unclear which are the best parameters to consider in making correlations. We therefore consider several possibly useful stellar parameters (bolometric luminosity, stellar mass, effective temperature, rotation, circumstellar disk properties, accretion rates) to look for relations to the X-ray emission level; note that relation between X-ray activity and age are discussed in a separate paper (Preibisch & Feigelson 2005). For the characterization of the X-ray properties we consider here the X-ray luminosity $`L_\mathrm{X}`$, the fractional X-ray luminosity $`L_\mathrm{X}/L_{\mathrm{bol}}`$, and the X-ray surface flux $`F_\mathrm{X}`$, i.e. X-ray luminosity divided by the stellar surface area. Throughout this paper, we use $`L_\mathrm{X}`$ to refer to the extinction-corrected total band ($`0.58`$ keV) luminosity $`L_{t,c}`$ defined and listed by Getman et al. (2005a). Most of the relations presented here were already studied in other X-ray data sets (e.g. Feigelson et al., 2003; Flaccomio et al., 2003b; Preibisch & Zinnecker, 2002), often with similar results to what we find here. Nevertheless, we study these relations here in some detail because our COUP data provide a unique, in fact the best data set for an investigation of the nature and the origin of the X-ray emission from TTS for the following reasons. With the exceptionally well characterized young stellar population in the ONC we can take advantage of known stellar parameters for several hundred stars. The high sensitivity of the COUP X-ray data yields a detection limit of $`L_{\mathrm{X},\mathrm{min}}10^{27.3}`$ erg/sec for lightly absorbed stars and allows us to detect more than 97% of the stars in the lightly absorbed optical sample of cluster members. Our analysis is therefore based on an nearly complete sample. The high sensitivity allows us to detect X-ray emission of the young solar-luminosity stars down to activity levels<sup>4</sup><sup>4</sup>4Note that the quoted activity level refers to the relatively hard 0.5–8 keV COUP band. This band covers most of the X-ray flux from TTS (which are characterized by rather high plasma temperatures of $`10`$ MK and have accordingly relatively hard X-ray spectra). Solar-like field stars and the Sun, however, exhibit considerably lower plasma temperatures ($`2`$ MK in the case of the Sun) and have accordingly softer X-ray spectra with most of the X-ray flux below the COUP band. If the Sun were located at the distance of the ONC, it would be only marginally detectable during its maximum phase of coronal activity. of $`\mathrm{log}\left(L_\mathrm{X}/L_{\mathrm{bol}}\right)6`$. Our statistical analysis strongly benefits from the large sample of 598 optically well characterized ONC stars for which X-ray emission is detected. This represents the largest homogenous sample of TTS that has ever been studied with very sensitive X-ray observations (and will remain so for the foreseeable future). The availability of X-ray spectra with good S/N for nearly all sources allows the determination of reliable X-ray luminosities. The 10 day long observation provides a much better measure for the “typical” X-ray properties of the strongly variable TTS than observations with shorter exposure times, which yield only a “snapshot”. ### 4.1 X-ray luminosity and bolometric luminosity The plot of X-ray luminosity versus bolometric luminosity is shown in Fig. 2. Nearly all stars show $`\mathrm{log}\left(L_\mathrm{X}/L_{\mathrm{bol}}\right)>5`$ and therefore are much more X-ray active than the Sun (for which $`\mathrm{log}\left(L_\mathrm{X}/L_{\mathrm{bol}}\right)6.5`$ is an average during the course of the solar cycle). The most active stars show fractional X-ray luminosities around $`\mathrm{log}\left(L_\mathrm{X}/L_{\mathrm{bol}}\right)=3`$, which is known as the “saturation limit” for coronally active stars (Fleming et al., 1995). Considering only the low-luminosity ($`L_{\mathrm{bol}}<10L_{}`$) stars, we find a clear correlation between X-ray and bolometric luminosity, although with a very large scatter. We utilized the ASURV survival analysis package (Feigelson & Nelson, 1985; Isobe et al., 1986; LaValley et al., 1990) for the statistical investigation of the relation between $`L_\mathrm{X}`$ and $`L_{\mathrm{bol}}`$. The ASURV software allows one to deal with data sets that contain non-detections (upper limits) as well as detections, and provides the maximum-likelihood estimator of the censored distribution, several two-sample tests, correlation tests and linear regressions. The linear regression fit with the parametric Estimation Maximization (EM) algorithm in ASURV yields $`\mathrm{log}\left(L_\mathrm{X}[\mathrm{erg}/\mathrm{sec}]\right)=30.00(\pm 0.04)+1.04(\pm 0.06)\times \mathrm{log}\left(L_{\mathrm{bol}}/L_{}\right)`$ with a standard deviation of 0.70 dex in $`\mathrm{log}L_\mathrm{X}`$ for the low-luminosity stars ($`L_{\mathrm{bol}}10L_{}`$). This relation is very similar to the relations found for other young clusters (cf. Feigelson & Montmerle 1999) and is consistent with a linear relation between X-ray and bolometric luminosity characterized by $`\mathrm{log}\left(L_\mathrm{X}/L_{\mathrm{bol}}\right)=3.6\pm 0.7`$. ### 4.2 X-ray activity and stellar mass Next we considered the relation between X-ray luminosity and stellar mass (Fig. 3). As described in Getman et al. (2005a), the stellar masses listed in the COUP tables were derived from the Siess, Dufour, & Forestini (2000, SDF hereafter) PMS models. It is well known that mass estimates from PMS evolutionary models are subject to significant uncertainties; different PMS models and/or temperature scales can lead to differences by as much as a factor of $`2`$ in the mass estimates (for detailed investigations of these uncertainties see, e.g., Luhman, 1999 or Hillenbrand & White, 2004). As a test to what extent the $`L_\mathrm{X}M`$ relation is dependent on the choice of the PMS model<sup>5</sup><sup>5</sup>5Our comparison here is restricted to the PMS models of Siess, Dufour, & Forestini (2000) and Palla & Stahler (1999) because these two models cover particularly wide ranges of stellar masses; this does not imply that we consider these specific models to be “better” than other models., we compared the relation found for the masses derived from the SDF models to those based on stellar masses estimated from the PMS models of Palla & Stahler (1999, PS hereafter). Note that the masses determined from these two sets of models agree very well with each other for objects with $`M>0.4M_{}`$, but below $`0.4M_{}`$ the PS models yield systematically lower masses than the SDF models. For both sets of stellar masses a clear correlation is found between X-ray luminosity and mass. For the low-mass ($`M2M_{}`$) stars, the SDF models lead to an EM linear regression fit of $`\mathrm{log}\left(L_\mathrm{X}[\mathrm{erg}/\mathrm{sec}]\right)=30.37(\pm 0.06)+1.44(\pm 0.10)\times \mathrm{log}\left(M/M_{}\right)`$ with a standard deviation of 0.65, whereas the PS models yield a somewhat shallower relation of $`\mathrm{log}\left(L_\mathrm{X}[\mathrm{erg}/\mathrm{sec}]\right)=30.34(\pm 0.05)+1.13(\pm 0.08)\times \mathrm{log}\left(M/M_{}\right)`$ with a standard deviation of 0.64. From this exercise we conclude that the detailed shape of the $`L_\mathrm{X}M`$ correlation does depend on the PMS model used, but the general dependence is independent of the choice of the model. The power-law slopes we find here for the ONC TTS are considerably lower than those found for the TTS in the Chamaeleon star forming region (slope $`=3.6\pm 0.6`$ in the mass range $`0.62M_{}`$; Feigelson et al., 1993) and the very young stellar cluster IC 348 (slope $`=2.0\pm 0.2`$ in the mass range $`0.12M_{}`$; Preibisch & Zinnecker, 2002) or than that derived for M-type field stars (slope $`=2.5\pm 0.5`$ in the mass range $`0.150.6M_{}`$; Fleming et al., 1988). The differences in the slopes are in part due to differences in the considered mass ranges and in the methods to estimate stellar masses. Another factor may be that many of the previous studies had to deal with large numbers of X-ray upper limits for undetected very-low mass stars, which perhaps caused the typical X-ray luminosities of these very-low mass stars to be underestimated. Next we consider the fractional X-ray luminosity as a function of mass. In Fig. 4 we show the $`L_\mathrm{X}/L_{\mathrm{bol}}M`$ relation for low-mass stars. The statistical tests in ASURV reveal a very shallow, but nevertheless highly significant ($`P(0)<10^4`$) correlation between fractional X-ray luminosity and mass for low-mass ($`M<2M_{}`$) objects. The linear regression fit with the EM algorithm yields $`\mathrm{log}\left(L_\mathrm{X}/L_{\mathrm{bol}}\right)=3.40(\pm 0.06)+0.42(\pm 0.11)\times \mathrm{log}\left(M/M_{}\right)`$ with a standard deviation of 0.69 dex. We also used this relation to illustrate the influence of X-ray variability during the COUP observation on the resulting correlations. If we consider the characteristic X-ray luminosities derived from the MLB lightcurve analysis rather than the average X-ray luminosities, we find a very similar relation with $`\mathrm{log}\left(L_{\mathrm{X},\mathrm{char}}/L_{\mathrm{bol}}\right)=3.56(\pm 0.05)+0.40(\pm 0.10)\times \mathrm{log}\left(M/M_{}\right)`$ with a standard deviation of 0.64 dex. This test shows that the use of the characteristic rather than the average X-ray luminosities decreases the values of the fractional X-ray luminosities slightly, but the power-law slopes for the $`L_\mathrm{X}/L_{\mathrm{bol}}M`$ correlations are nearly identical, and the scatter in the correlation diagrams is only very slightly smaller. From this exercise, we conclude that the correlations do not significantly depend on the choice of the average or characteristic X-ray luminosities. ### 4.3 X-ray activity and stellar effective temperature In Fig. 5 we show the X-ray surface fluxes (i.e. X-ray luminosities divided by the stellar surface area) of the ONC TTS plotted versus their effective temperatures. Nearly all stars have X-ray surface fluxes in the range $`10^410^8\mathrm{erg}/\mathrm{cm}^2/\mathrm{sec}`$, which corresponds nicely to the minimum and maximum X-ray surface flux found for different structures in the solar corona (where coronal holes and the background corona show X-ray fluxes around $`10^4\mathrm{erg}/\mathrm{cm}^2/\mathrm{sec}`$, while active regions show fluxes up to $`10^8\mathrm{erg}/\mathrm{cm}^2/\mathrm{sec}`$); the similarity of the X-ray surface flux ranges found for late-type stars and for different constituents of the solar corona has already been noted, e.g., in Schmitt (1997) or Peres et al. (2004). The plot also shows a strong decline of the X-ray surface flux with effective temperature among the M-type stars. This dependence is much more pronounced than the very shallow relation between $`L_\mathrm{X}/L_{\mathrm{bol}}`$ and the stellar mass discussed above. This effect can be understood if one recalls that the surface flux and the fractional X-ray luminosity are related to each other by $`F_\mathrm{X}T_{\mathrm{eff}}^4\times \left(L_\mathrm{X}/L_{\mathrm{bol}}\right)`$. Therefore, $`F_\mathrm{X}`$ decreases with decreasing $`T_{\mathrm{eff}}`$ for constant $`L_\mathrm{X}/L_{\mathrm{bol}}`$. ### 4.4 Comparison to main-sequence stars For a meaningful comparison of the ONC TTS to main-sequence (MS) stars, it is important to keep in mind that our ONC TTS sample is an optically selected and representative sample of cluster members. For a proper comparison we therefore have to use an optically selected (not X-ray selected) sample of MS stars. A well suited comparison sample is the NEXXUS database (Schmitt & Liefke, 2004), which provides updated ROSAT X-ray and optical data (including accurate HIPPARCOS parallaxes) for nearby field stars. It contains volume-limited samples for G-type ($`d_{\mathrm{lim}}=14`$ pc), K-type ($`d_{\mathrm{lim}}=12`$ pc), and M-type ($`d_{\mathrm{lim}}=6`$ pc) stars with detection rates of more than 90%. The NEXXUS tables were kindly provided to us by the authors; they list $`M_V`$, $`BV`$, $`L_\mathrm{X}`$ in the $`0.12.4`$ keV ROSAT band, and the X-ray surface flux $`F_\mathrm{X}`$. We used the data for the 43 G-type stars (including 4 non-detections), the 54 K-type stars (including 2 non-detections), and the 79 M-type stars (including 5 non-detections). Bolometric luminosities, effective temperatures, and masses of the stars were estimated by interpolation using MS relationships of these quantities with the absolute magnitude $`M_V`$. When comparing the NEXXUS data to our COUP data one has to take into account the different energy bands for which X-ray luminosities were computed. The NEXXUS X-ray luminosities are given for the $`0.12.4`$ keV ROSAT band, and for comparison with our COUP results we have to transform these luminosities into $`0.58`$ keV band. The transformation factor can be calculated with PIMMS and depends on the X-ray spectrum; since we can assume thermal plasma spectra, the transformation mainly depends on the plasma temperature. For the NEXXUS stars, the count-rate to luminosity transformation factor used by Schmitt & Liefke (2004) assumes a plasma temperature of $`2.5`$ MK, and the corresponding energy band correction factor is 0.33 dex. In the following comparisons we also consider the X-ray properties of our Sun. For this, we use here the ROSAT-band X-ray luminosity range of $`\mathrm{log}\left(L_\mathrm{X}[\mathrm{erg}/\mathrm{sec}]\right)=26.827.9`$ based on the results of Judge et al. (2003) for the activity range of a typical solar cycle. Assuming a plasma temperature of 2 MK, the flux in the $`0.58`$ keV band is 0.48 dex lower than that in the $`0.12.4`$ keV band. Figure 6 compares the $`L_\mathrm{X}L_{\mathrm{bol}}`$ relations for the COUP optical sample to that for the NEXXUS sample of field stars. A clear correlation between X-ray and bolometric luminosity is not only seen for the ONC TTS, but also for the NEXXUS field stars. The tests in ASURV show that the correlation for the NEXXUS stars is significant $`(P(0)<10^4)`$; the linear regression fit with the EM algorithm for the G-, K-, and M-type stars in the NEXXUS sample yields a power-law slope of $`0.42\pm 0.05`$, which is much shallower than the slope found for the $`L_\mathrm{X}L_{\mathrm{bol}}`$ correlation for the COUP stars ($`1.04\pm 0.06`$). Figure 7 compares the $`L_\mathrm{X}M`$ relations for the COUP optical sample to that for the NEXXUS sample of field stars. It is interesting to see that there is a clear correlation between X-ray luminosity and mass for the NEXXUS field stars. The tests in ASURV show the X-ray luminosity and mass are clearly correlated $`(P(0)<10^4)`$; the linear regression fit with the EM algorithm for objects in the mass range $`0.082M_{}`$ yields $`\mathrm{log}\left(L_\mathrm{X}[\mathrm{erg}/\mathrm{sec}]\right)=27.58(\pm 0.07)+1.25(\pm 0.15)\times \mathrm{log}\left(M/M_{}\right)`$ with a standard deviation of 0.77. The corresponding correlation for the TTS in the COUP optical sample yielded a power-law slope of $`1.44\pm 0.10`$, which is consistent to the slope for the field stars within the uncertainties. The similarity of the slopes found in the $`L_\mathrm{X}M`$ relations for the ONC TTS and the field stars may indicate that the relation between stellar mass and X-ray luminosity is more fundamental than that between bolometric and X-ray luminosity. The plot of fractional X-ray luminosities against stellar masses (Fig. 7) shows that some of the very-low mass field stars reach similar activity levels as the TTS. The solar-mass field stars, on the other hand, are typically much less X-ray active than their young COUP counterparts. The different activity levels of the field stars as a function of mass can be understood as a consequence of the activity-rotation relation for MS stars: many of the very-low mass field stars are rapid rotators, thus show high levels of X-ray activity, while most solar-mass field stars rotate quite slowly, therefore displaying lower activity levels. For the COUP stars, on the other hand, we show in Section 5 that all stars with known rotation period rotate more rapidly than the Sun, and that their X-ray activity is probably unrelated to their rotation period. ## 5 X-ray emission, rotation, and convection ### 5.1 The activity–rotation relation for the TTS For MS stars, the well established correlation between fractional X-ray luminosity and rotation period (e.g. Pallavicini et al., 1981; Pizzolato et al., 2003) constitutes the main argument for solar-like dynamo mechanism as the origin of their X-ray activity. As already noted in the introduction, the existence of a similar activity – rotation relation could not be unambiguously proven for PMS stars, mainly due to a lack of statistical power in the underlying data (in most studies the sample sizes were too small for statistically significant conclusions to be drawn). The previous $`Chandra`$ ONC studies (Feigelson et al., 2002a; Flaccomio et al., 2003b), however, provided strong evidence that the TTS do not follow the activity – rotation relation for MS stars. Table 9 in Getman et al. (2005a) lists rotation periods for 158 stars in our COUP optical sample. Considering also the additional rotation data as listed in Flaccomio et al. (2005), rotation periods from photometric monitoring are available for 228 stars in our optical TTS sample (169 M-type stars, 58 K-type stars, and one G-type star). In Fig. 8 we plot the fractional X-ray luminosity versus rotation period for these stars, and compare them to data for MS stars. It is rather obvious that the COUP stars do not follow the well established activity-rotation relation shown by the MS stars, i.e. increasing activity for decreasing rotation periods followed by saturation at $`L_\mathrm{X}/L_{\mathrm{bol}}10^3`$ for the fastest rotators. A statistical analysis reveals for the COUP stars a correlation between $`L_\mathrm{X}/L_{\mathrm{bol}}`$ and $`P_{\mathrm{rot}}`$ rather than the anti-correlation seen for the MS stars: the linear regression analysis with SLOPES (Isobe et al., 1990; Feigelson & Babu, 1992) yields a bisector regression fit of the form $`\mathrm{log}\left(L_\mathrm{X}/L_{\mathrm{bol}}\right)=4.21(\pm 0.07)+1.27(\pm 0.09)\times \mathrm{log}\left(P_{\mathrm{rot}}[\mathrm{days}]\right)`$. This correlation is statistically significant; a Kendall’s $`\tau `$ and Spearman’s $`\rho `$ test give probabilities of $`P(0)=0.0002`$ for the null hypothesis that a correlation is not present. The observed correlation between $`L_\mathrm{X}/L_{\mathrm{bol}}`$ and $`P_{\mathrm{rot}}`$ is clearly very different from the anti-correlation shown by the MS stars (where the bisector regression fit yields a slope of $`2.35\pm 0.16`$ for the sample shown in Fig. 8 in the period range 1–10 days). These results do not change significantly if we use the characteristic rather than the average X-ray luminosities. Before we consider possible explanations for these findings, it is important to note that rotation periods are known for only $`38\%`$ of the X-ray detected stars in our optical sample, and that this subsample may be biased with respect to its X-ray properties. Stassun et al. (2004a) studied archival ACIS data of the ONC and pointed out that the stars with known rotation period in their sample show systematically higher X-ray activity than the stars with unknown periods. A similar difference is present in our COUP optical sample: the median fractional X-ray luminosity for the TTS with known rotation periods is at $`\mathrm{log}\left(L_\mathrm{X}/L_{\mathrm{bol}}\right)=3.31`$, while the median value for TTS with unknown periods is $`3.71`$. A KS test gives a probability of $`P(0)10^4`$ for the hypothesis that the distributions of fractional X-ray luminosities in both samples are identical, i.e. the apparent difference of about a factor of $`2.5`$ in X-ray activity is statistically significant. This difference is not due to systematic differences in the basic stellar parameters<sup>6</sup><sup>6</sup>6We note that the stars with known periods have systematically higher masses ($`\mathrm{\Delta }\mathrm{log}\left(M\right)0.06`$ dex, $`P_{\mathrm{KS}}(0)=0.006`$) than the stars without periods. According to the correlation between stellar mass and fractional X-ray luminosity established in Sect. 4, this difference in stellar masses would, however, predict a difference in the activity level of only $`\mathrm{\Delta }\mathrm{log}\left(L_\mathrm{X}/L_{\mathrm{bol}}\right)0.02`$ dex, much smaller than the observed $`0.4`$ dex difference between stars with and without known periods. This effect can therefore not explain the difference in the activity levels of stars with and without known periods. of the two samples. The explanation of this difference was discussed in detail by Stassun et al. (2004a): Rotation periods can only be determined for stars showing sufficiently large spot related photometric variability. The level of photometric variability, however, is related to the level of magnetic activity, and therefore the more active stars (i.e. those with higher X-ray luminosities) are easier targets for a determination of photometric rotation periods, while the less active stars (i.e. those with lower X-ray luminosities) show too small photometric variability to allow determination of periods. A quantitative description of these interrelations for the case of MS stars has been given by Messina et al. (2003). It is therefore likely that the COUP stars without known rotation periods rotate on average slower than the stars with known periods. This introduces a bias our sample, since most of the missing stars (i.e. those without known rotation periods) have lower fractional X-ray luminosities and longer rotation periods than the stars in our sample. The apparent correlation between X-ray activity and rotation period may therefore (in part) be due to these selection effects. Nonetheless, it appears very unlikely that the ONC TTS follow the same rotation-activity relation as found for MS stars. First, we note that the effects of the bias due to unknown rotation periods appear much too small to transform the strong $`L_\mathrm{X}/L_{\mathrm{bol}}P_{\mathrm{rot}}`$ anti-correlation of the MS stars into an apparent positive correlation. Second, the quite high fractional X-ray luminosities for most TTS in the period range between $`7`$ and $`15`$ days, which are more than one order of magnitude higher than the typical values for MS stars, also indicate differences between the TTS and the MS sample. In conclusion, we are confident that the ONC TTS do not follow the same rotation-activity relation as seen in MS stars, but it remains unclear whether and how the X-ray activity of the TTS is correlated to their rotation periods. The inability to draw meaningful conclusions from our X-ray and rotation data may be due to the fact that the rotation period is perhaps the wrong variable to look at. Theoretical studies of the solar-like $`\alpha \mathrm{\Omega }`$ dynamos show that the dynamo number is not directly related to the rotation period, but to more complicated quantities such as the radial gradient of the angular velocity and the characteristic scale length of convection at the base of the convection zone. It can be shown that (with some reasonable assumptions) the dynamo number is essentially proportional to the inverse square of the Rossby number $`Ro`$ (e.g. Maggio et al., 1987). The Rossby number is defined as the ratio of the rotation period to the convective turnover time $`\tau _c`$, i.e. $`Ro:=P_{\mathrm{rot}}/\tau _c`$. For MS stars, the theoretical expectations are well confirmed: it is well established (e.g. Montesinos et al., 2001) that the stellar activity shows a tighter relationship to the Rossby number than to rotation period. The shape of the relation is similar to that of the activity–rotation relation: for large Rossby numbers, activity rises strongly as $`L_\mathrm{X}/L_{\mathrm{bol}}Ro^2`$ until saturation at $`L_\mathrm{X}/L_{\mathrm{bol}}10^3`$ is reached around $`Ro0.1`$, which is followed by a regime of “supersaturation” for very small Rossby numbers, $`Ro0.02`$. The convective turnover time scale is a sensitive function of the physical properties in the stellar interior and its determination requires detailed stellar structure models. Many studies of stellar activity therefore used semi-empirical interpolations of $`\tau _c`$ values as a function of, e.g., $`BV`$ color. This may be appropriate for MS stars, but seems to be insufficient for TTS which have a very different and quickly evolving internal structure. ### 5.2 Computation of convective turnover times for the TTS For the computation of convective turnover times a series of stellar evolution models with masses ranging from $`0.065M_{}`$ to $`4.0M_{}`$ was computed with the Yale Stellar Evolution Code. The evolution was assumed to start at the stellar birthline, where stars initially become visible objects (Palla & Stahler, 1993). All models used the parameters derived for the standard solar model, where the initial $`X`$, $`Z`$, and the mixing-length ratio were varied until a solar model at the solar age of 4.55 Gyr has the observed solar values of luminosity, radius, and $`Z/X`$ (=0.0244; Grevesse et al., 1996). The model that best matched the solar properties<sup>7</sup><sup>7</sup>7We note that new precision measurements of elemental abundances on the solar surface imply a lower metalicity than previously assumed, and lead to inconsistencies in theoretical solar models with respect to the depth of the solar convection zone (see Bahcall et al., 2004, and reference therein). was of $`(X,Z)_{\mathrm{initial}}=(0.7149,0.0181)`$ and the mixing length ratio 1.7432. These values were then used for all stellar models. A detailed discussion of the physics used in this study for the construction of stellar models can be found in Yi et al. (2001). The most important aspects are as follows: The solar mixture was assumed as given by Grevesse & Noels (1993). For $`\mathrm{log}T>4`$ OPAL Rosseland mean opacities (Iglesias & Rogers, 1996) were used, for $`\mathrm{log}T<4`$ opacities from Alexander & Ferguson (1994). The OPAL EOS 2001 equation of state (Rogers et al., 1996) was used and the energy generation rates were set according to Bahcall & Pinonneault (1992). Neutrino losses were taken following Itoh et al. (1989), and for the helium diffusion the values of Thoul et al. (1994) were used. Since the dynamo action is believed to take place at the base of the convection zone, anchored in the radiative layers just below the convective interface, the convective turnover time of the deepest part of the convection zone is the most relevant in the evaluation of the Rossby number. Our knowledge of stellar convection is too limited to calculate ’correct’ convection turnover times, because the characteristic length scales, as well as the velocities, are not well known. Even when one decides to resort to the mixing length approximation, there are still uncertainties: the mixing length ratio is assumed to be the same for all stars with different masses and/or at different evolutionary stages, which is probably not fully correct. However, for the convection near the base of the convection zone where the temperature gradient is for all practical purposes adiabatic, the mixing-length approximation is known to provide an adequate description of convection at least in an average sense (Kim & Demarque, 1996). For the characteristic timescale of convective overturn, the convective turnover time (i.e. the local mixing length divided by the local velocity) was calculated at each time step, which was determined at a distance of one-half the mixing length above the base of the surface convection zone<sup>8</sup><sup>8</sup>8Note that for fully convective stars the base of the convection zone is the center. (Gilliland, 1986; Kim & Demarque, 1996). The convective turnover times were determined for the stars in the optical sample according to the model that represents their corresponding state in the HR-diagram, i.e. reproduces the $`(T_{\mathrm{eff}},L_{\mathrm{bol}})`$ values. The resulting values are shown in Fig. 9. Note that the convective turnover times of the TTS are much larger (up to factors of $`8`$) than those in MS stars, and depend strongly on the effective temperatures and ages of the stars. ### 5.3 The activity – Rossby number relation for the TTS The Rossby numbers for the ONC TTS were computed by dividing their rotation periods by the values for their local convective turnover time as derived above. The plot of fractional X-ray luminosity versus Rossby number in Fig. 10 shows no strong relation between these two quantities. All TTS have Rossby numbers $`<0.2`$ and therefore are in the saturated or super-saturated regime of the activity – Rossby number relation for MS stars. To search for indications of super-saturation, we compared the fractional X-ray luminosities of the TTS in the saturated ($`0.1>Ro>0.02`$) and super-saturated ($`Ro0.02`$) regimes. Indeed, we found a slightly lower median $`\mathrm{log}\left(L_\mathrm{X}/L_{\mathrm{bol}}\right)`$ of $`3.63`$ for the TTS in the super-saturated regime as compared to $`3.43`$ for those in the saturated regime; a KS test gives a probability of $`P(0)=0.031`$ for the assumption of equal $`L_\mathrm{X}/L_{\mathrm{bol}}`$ distributions in both samples, i.e. the difference is significant at the 97% level. Thus, the fractional X-ray luminosities of the TTS show a qualitatively similar relation to their Rossby numbers as is found for MS stars. A remarkable difference between our TTS sample and the data for MS stars is the very wide dispersion of fractional X-ray luminosities at a given Rossby number in our TTS data. The scatter extends over about three orders of magnitude, and even if we consider the characteristic (i.e. flare-cleaned) rather than the average X-ray luminosities, the scatter is only very slightly smaller. This large scatter is in strong contrast to the tight relation found for MS stars, where the scatter in $`\mathrm{log}\left(L_\mathrm{X}/L_{\mathrm{bol}}\right)`$ at a given Rossby number is only about $`\pm 0.5`$ dex (e.g., Pizzolato et al., 2003). To conclude, we find that the X-ray activity of the ONC TTS may depend on Rossby numbers in a similar way to what is found for MS stars, but the large scatter of X-ray activity at any given Rossby number suggests that additional factors are important for the level of X-ray activity. ## 6 Possible explanations for the wide scatter in the correlations All the correlations between the X-ray activity and other stellar parameters show a very large scatter, often exceeding three orders of magnitudes. Three obvious reasons for the presence of scatter are uncertainties in the variables, X-ray variability, and unresolved binaries. Can these effects account for the wide scatter seen in the correlations? We first consider X-ray variability. Most of the COUP sources show strong variability in their lightcurves. It is well known that variability is a common feature in the X-ray emission from TTS, active stars, and also our Sun. The degree of variability is a function of the timescales. For active MS stars and the Sun, short term variability (timescales of minutes to hours) is usually dominated by flares which can cause large variations (sometimes exceeding factors of 10). On timescales from days to weeks, the typical variations are about a factor of 3 and up to 10, while the typical variability on timescales from months to years is a factor of about $`34`$ (e.g. Micela & Marino, 2003; Orlando et al., 2004). Furthermore, indications for stellar X-ray activity cycles have been found for some stars; these cycles can induce more than one order of magnitude variability over a few years (e.g., Favata et al., 2004). In previous, generally much shorter X-ray observations of TTS, the large scatter was often assumed as being probably due in part to X-ray flaring. The COUP data provide two important pieces of information in this respect. First, the influence of individual flares on the average X-ray luminosity is strongly smoothed out by the long time basis of the COUP observation; the scatter seen in the correlations based on our COUP data should therefore be much smaller than what would be found from shorter, snapshot-like observations. Second, the use of the characteristic X-ray luminosities, which effectively exclude periods of flaring from the lightcurves, should further reduce the scatter, if short-term X-ray variability were the main reason for the large scatter in the correlation diagrams. These expectations are, however, not confirmed in our data: We find that the scatter in the COUP (i.e., 10 day average) X-ray luminosities of the ONC TTS, e.g. as a function of bolometric luminosity or mass, is very similar to that found in the correlations based on the previous 23 hr ACIS observation (Feigelson et al., 2002a). Also, the use of characteristic rather than average X-ray luminosities reduces the scatter only marginally. This suggests that variability on time-scales between $`1`$ and $`10`$ days is not the main source of the large scatter in the correlations between X-ray activity and other stellar parameters. What about variability on longer timescales? We can investigate the variations on timescales of several years by comparison of the COUP data (obtained in January 2003) to the previous 23 hour ACIS observations of the ONC (obtained in October 1999/April 2000) given by Feigelson et al. (2002a). For 515 of the COUP detected stars in our optical sample X-ray luminosities based on the 23 hour observations are listed by Feigelson et al. (2002a). Since the spectral fitting procedure used in the analysis of the 23 hr data is not identical to that used for the COUP data, we compare here the observed X-ray luminosities, $`L_\mathrm{t}`$, without extinction correction, which just give the integral of the observed flux over the spectrum. We find a good agreement of the luminosities from the two observations separated by more than 3 years: the median absolute deviation from equal luminosities is only 0.31 dex, corresponding to just a factor of $`2`$. This demonstrates that the X-ray luminosities of most ONC TTS vary only slightly on timescales of several years. The observed level of variability cannot account for the large scatter seen in the correlation diagrams. Now we try to quantify the uncertainties of the variables in the correlation diagrams. According to H97, the uncertainties in $`\mathrm{log}\left(L_{\mathrm{bol}}\right)`$ are $`0.2`$ dex. The uncertainties in the stellar mass estimates are probably of a similar magnitude. The uncertainties of the X-ray luminosities, derived from the spectral fits, are difficult to determine because the spectral models are highly non-linear and the individual spectral parameters are often strongly correlated. Furthermore, ambiguities can occur when two qualitatively different spectral models give similarly good fits. We therefore assume the typical random uncertainties of $`\mathrm{log}\left(L_\mathrm{X}\right)`$ to be similar to the uncertainties in the emission measures derived in the spectral fits, i.e. about $`0.15`$ dex. Note, however, that some sources may be affected by systematic errors, which may well exceed this level. Finally, we consider the effect of unresolved binaries. The presence of an unresolved companion causes an overestimation of both, the bolometric and the X-ray luminosity. However, if X-ray and bolometric luminosity are related linearly (as our data suggest), unresolved companions should only shift the position of a star in the $`L_\mathrm{X}L_{\mathrm{bol}}`$ diagram along the correlation line and not increase the scatter. The stellar mass determined by comparison with PMS tracks depends (in the case of low-mass stars) mainly on the measured spectral type; as the combined optical spectrum of a binary system is dominated by the light of the primary component, the inferred mass is that of the primary, whereas the observed X-ray luminosity is the sum for primary and companion. If X-ray luminosity and stellar mass are correlated, the overestimation of the X-ray luminosity should be at most a factor of two<sup>9</sup><sup>9</sup>9There may, however, be larger effects in special cases. For example, in the case of a binary system in which both components show very different amounts of extinction, a fit to the combined X-ray spectrum may easily lead to wrong parameters.; the typical shift in $`\mathrm{log}\left(L_\mathrm{X}\right)`$ depends on the distribution of mass ratios in the binary systems (which is not well known), but is presumably about $`0.20.3`$ dex among the low-mass stars. The combined effect of the variability, the uncertainties in the variables, and unresolved binaries should therefore produce a scatter of roughly $`\pm (0.40.5)`$ dex. This is considerably less (by about a factor of two) than the scatter observed in the correlation diagrams, where the standard deviations are typically $`0.7`$ dex. A large fraction of the scatter in the correlations must therefore be due to other reasons, most likely due to intrinsic differences in the X-ray activity levels of the TTS. In §8 we will show that the large scatter is probably related to the influence of accretion on the X-ray properties. ## 7 X-ray plasma temperatures As described in G05, the X-ray spectra of the COUP sources were fitted with single-temperature or (in most cases) two-temperature thermal plasma models plus absorption. We are fully aware that these relatively simple models are not “perfect”, since it is well known that the coronae of active stars are generally not monothermal and can usually not be considered as to consist of just two different temperature components (e.g. Brickhouse et al., 2000; Sanz-Forcada et al., 2003). Nevertheless, this approach is justified because the purpose of our analysis was to characterize the coronal temperatures of the COUP stars in a homogenous way using a parsimonious model, rather than to perform a detailed investigation of the temperature structure of those sources (which will be the topic of separate studies). As demonstrated for example by Peres et al. (2000), fits of simulated spectra based on continuous temperature distributions with simple one- or two-temperature models usually yield temperatures near the peaks of the underlying temperature distribution. We therefore assume that the temperatures derived from the fits reflect some kind of a “characteristic” temperature, which then can be related to other stellar parameters. In Fig. 11 we plot the plasma temperatures and the ratios of the emission measures of the hot and cool component derived in the spectral fits versus the X-ray surface flux. First, we note that the temperature of the hot plasma component increases with increasing surface flux; the slope is consistent the relation $`F_XT^6`$, what is similar to a scaling relation derived for the solar corona by Peres et al. (2004) (which however, was established for a much lower temperature range than seen here on the TTS). Second, we note that the relative contribution from the hot component (as measured by the ratio of emission measures for the hot and cool components) also increases with increasing X-ray activity. The most interesting result from these plots is the remarkable similarity of the temperatures of the cool plasma component in our sample. For nearly all stars a temperature of about 10 MK is found for the cool component. This suggests that this 10 MK component is a real feature in the coronal temperature distribution of the TTS. It is interesting to note that a $`10`$ MK plasma component seems to be some kind of a general feature of coronally active stars; for example, Sanz-Forcada et al. (2003) determined the emission measure distribution of 28 coronally active nearby field stars and found a peak at 8–10 MK in most of their stars. They argued that this temperature component may define a fundamental coronal structure, which is probably related to a class of compact loops with high plasma density. Figure 12 shows the derived plasma temperatures as a function of stellar effective temperature. Among the M-type stars one can see a decrease in the temperature of the hot component for decreasing effective temperature. This can be understood as a consequence of the correlation of the hot plasma temperature with the level of X-ray activity (as traced by $`F_\mathrm{X}`$ or $`L_\mathrm{X}/L_{\mathrm{bol}}`$) and the decrease of X-ray activity with decreasing mass or effective temperature (see, e.g. Fig. 5). In Fig. 13 we plot the temperature of the hot versus that of the cool plasma component for the ONC TTS. We also have included temperatures derived for G- and K-type stars in several young clusters and for solar-like field stars, as well as values for different structures in the solar corona as derived in the simulations of the “Sun as a star” by Orlando et al. (2004). The ONC TTS follow the general correlation between the temperatures of the hot and cool plasma components seen for the MS stars, although nearly all TTS show much higher temperatures than found on the MS stars. In the solar corona and in many active MS stars, plasma temperatures significantly exceeding 10 MK are only seen during strong flares. The high plasma temperatures found for the TTS may suggest an increased contribution of flares to the total X-ray emission. ## 8 X-ray emission and circumstellar accretion disks It is still unclear how the (magnetic) interactions between a young stellar object and its surrounding circumstellar accretion disk influence the X-ray activity. Many X-ray observations of star-forming regions have been used to search for differences in the X-ray properties of the classical T Tauri stars (CTTS; usually defined by the presence of H$`\alpha `$ emission with equivalent widths $`W(\mathrm{H}\alpha )10`$ Å), which are thought to be actively accreting via circumstellar disks, and the weak-line T Tauri stars (WTTS; $`W(\mathrm{H}\alpha )<10`$ Å), which seem to lack disks and active accretion. The results of these investigations showed confusing differences: some studies (e.g. Gagné et al. 1995; Feigelson et al. 1993; Casanova et al. 1995; Preibisch & Zinnecker 2002) found no significant differences between the X-ray luminosity functions of CTTS and WTTS, while other studies, most notably the Taurus-Auriga study by Stelzer & Neuhäuser (2001), found clear differences in the X-ray luminosity functions, with the WTTS being the stronger X-ray emitters. The two $`Chandra`$ studies of the ONC before COUP also yielded seemingly contradictory results: Feigelson et al. (2002a) found no differences in the X-ray activity levels of stars with and without disks, whereas Flaccomio et al. (2003b) reported a strong difference in the X-ray activity levels of accreting and non-accreting stars. It is important to note that the different studies used different criteria to define the samples of CTTS and WTTS: sometimes the strength of the H$`\alpha `$ or Ca emission line were used, while other studies used the presence or absence of near-infrared excess emission. These different kinds of indicators actually measure different things that cannot be directly compared: H$`\alpha `$ or Ca line emission is thought to be a tracer of accretion, while infrared excess emission is a tracer of circumstellar material. While accretion obviously requires the presence of circumstellar material, the presence of circumstellar material alone does not necessarily mean that the young stellar object is also accreting. Furthermore, this issue is easily affected by strong selection effects if the samples of TTS are either X-ray selected or based on optical selection criteria such as emission lines or infrared excess. For example, in many star forming regions H$`\alpha `$ emission was used as a tracer to find and define the population of T Tauri stars. This can easily introduce a strong bias, because the CTTS are quite easy to recognize by their prominent H$`\alpha `$ emission even if they are very faint, whereas WTTS of similar brightness are much harder to identify. The optical H$`\alpha `$ selected samples of TTS are therefore often very incomplete for WTTS and much more complete for the CTTS (see discussion in Preibisch & Zinnecker, 2002). In fact, the majority of WTTS in many star forming regions have been found through X-ray observations (cf. Neuhäuser 1997) and therefore suffer from an X-ray selection bias. The COUP study provides us with the important advantage that we can use a statistically complete sample of all optically visible stars in the region, which does not suffer from any of the selection effects described above. In the following we will investigate how and in which way the X-ray activity is related to infrared excess emission as a tracer of circumstellar material (§8.1), optical line emission as a tracer of accretion (§8.2), and estimates of accretion rates and luminosities from astrophysical models (§8.3). ### 8.1 X-ray activity and infrared excess (= inner disk tracer) A good way to discern between TTS with and without circumstellar material is to look for infrared excess emission. The COUP tables list the color excess $`\mathrm{\Delta }\left(IK\right)`$ as determined by Hillenbrand et al. (1998). This quantity represents the color excess relative to the expected photospheric colors for the star’s spectral type after correction for reddening due to extinction, and has been shown to be a useful tracer of circumstellar material (see discussion in Hillenbrand et al., 1998). However, the $`\mathrm{\Delta }\left(IK\right)`$ excess is not optimal for detecting circumstellar material, since the $`K`$-band excess traces only the hottest dust in the innermost regions near the central star; it has been shown (e.g. Haisch et al. 2001) that many stars with circumstellar material show significant excess emission only at longer wavelengths. We therefore also used the $`L`$-band photometric data listed in the COUP tables and determined a color excess $`\mathrm{\Delta }\left(KL\right)`$ in an analog way as $`\mathrm{\Delta }\left(IK\right)`$. The $`\mathrm{\Delta }\left(KL\right)`$ excess has the advantage of being more sensitive for circumstellar material and less strongly affected by uncertainties in the visual extinction, but the disadvantage of being available for fewer (228) TTS than $`\mathrm{\Delta }\left(IK\right)`$ (446 TTS) in the optical sample. Since the infrared excess emission in our sample is correlated to effective temperature and stellar mass, we compare in Fig.14 and Fig. 15 stars with and without excesses in mass stratified samples. In the samples based on $`\mathrm{\Delta }(IK)`$ excess, we find significantly different ($`P(0)<0.02`$) fractional X-ray luminosity distributions for the $`0.10.2M_{}`$ and the $`0.20.3M_{}`$ bins, whereas for the more massive stars the differences are only of marginal statistical significance. If we use the $`\mathrm{\Delta }(KL)`$ excess (which we regard as more reliable), no significant differences are found for any of the mass ranges considered. ### 8.2 X-ray activity and Ca II line emission (= accretion tracer) Next we consider the presence/absence of signs for active accretion rather than disks. We follow the strategy of Hillenbrand et al. (1998), who used the equivalent width of the $`8542`$ Å Ca II line as an indicator of disk accretion. As they noted, stars without or with very weak accretion are expected to show this line in absorption with $`W(\mathrm{Ca}\mathrm{II})3`$ Å and a rather weak dependence on the spectral type. In more strongly accreting stars, the line gets filled and the equivalent width should be related to the mass accretion rate. In their analysis of the Chandra HRC data of the ONC, Flaccomio et al. (2003) classified stars as strong accretors if their Ca II line is in emission and has an equivalent width of $`W(\mathrm{Ca}\mathrm{II})<1`$ Å, while stars with the Ca II line in absorption with $`W(\mathrm{Ca}\mathrm{II})>1`$ Å are assumed to be not (or at most very weakly) accreting. Using this scheme, 142 (136) of the objects in our (lightly absorbed) optical sample are classified as strong accretors, and 134 (123) objects as weak or non-accretors. In the following text, we will simply denote these two groups as “accretors” and “non-accretors”. Flaccomio et al. (2003) found a clear difference in the (fractional) X-ray luminosities of the accretors and non-accretors in their $`Chandra`$ HRC data of the ONC, with the accretors being considerably (about a factor of 2–3) less X-ray bright than the non-accretors. A similar, although less pronounced, effect is found in our COUP data. In Fig. 16 we show the distributions of fractional X-ray luminosities of accreting and non-accreting stars in different mass bins. A significant difference between accretors and non-accretors is only found for the $`0.30.5M_{}`$ stars, whereas the other mass ranges show only marginal or no evidence at all for a difference in the distribution functions of fractional X-ray luminosities. To investigate the difference between accreting and non-accreting stars further, we compare in Fig. 17 the correlations between characteristic X-ray luminosities and the bolometric luminosities for the non-accretors and the accretors. The non-accretors show a very well defined correlation between $`L_{\mathrm{X},\mathrm{char}}`$ and $`L_{\mathrm{bol}}`$, the linear regression fit with the EM algorithm gives a power-law slope $`1.1\pm 0.1`$ and standard deviation of $`0.52`$ dex in $`\mathrm{log}\left(L_{\mathrm{X},\mathrm{char}}\right)`$. For the accretors, the correlation is much less well defined; the linear regression fit with the EM algorithm gives a power-law slope $`0.6\pm 0.1`$ and standard deviation of $`0.72`$ dex in $`\mathrm{log}\left(L_{\mathrm{X},\mathrm{char}}\right)`$. Very similar results are found for the correlations between characteristic X-ray luminosities and stellar masses for the non-accretors and the accretors. As discussed in §6, the scatter in the correlations due to X-ray variability, the uncertainties in the variables, and unresolved binaries is expected to be about $`0.40.5`$ dex. The standard deviation found for the $`L_\mathrm{X}L_{\mathrm{bol}}`$ correlation (or the $`L_\mathrm{X}M`$ correlation) for the non-accreting stars agrees well to that expectation, whereas the accreting stars show a considerably larger scatter. Some fraction of this scatter may be due to the fact that the more rapidly accreting stars may have larger errors compared to non-accreting stars in stellar luminosity and effective temperature values due to the effects of accretion on the observables that lead to these quantities. Another important result is found when considering the mean fractional X-ray luminosities: For the non-accretors, the median value for $`\mathrm{log}\left(L_\mathrm{X}/L_{\mathrm{bol}}\right)`$ is $`3.31`$, what is in very good agreement to the mean saturation value for rapidly rotating low-mass ($`0.220.6M_{}`$) field stars derived by Pizzolato et al. (2003). This means that the fractional X-ray luminosities of the non-accreting TTS are consistent to those of much older coronally active field stars. For the accretors, on the other hand, we find a median value for $`\mathrm{log}\left(L_\mathrm{X}/L_{\mathrm{bol}}\right)`$ of $`3.74`$, which is a factor of about 3 lower than the saturation value for fast rotating low-mass field stars. The accreting TTS therefore are responsible for the “X-ray deficit” of the ONC TTS, i.e. the fact that the median fractional X-ray luminosity of the TTS is lower than that found for rapidly rotating MS stars. We also looked for possible relations between the X-ray activity and the rotation rates of the accretors and non-accretors. The rotation periods found for the non-accretors and accretors overlap strongly, but the non-accretors show shorter median periods ($`3.8`$ days) than the accretors ($`6.8`$ days); a KS test gives a probability of only 0.03 that the distribution of periods is identical in both groups. However, neither for the accretors nor for the non-accretors statistically significant correlations are found between activity and rotation. The median fractional X-ray luminosities of the non-accretors and accretors with known rotation period are nearly identical, $`3.3\pm 0.4`$ and $`3.4\pm 0.8`$. The difference in X-ray activity between accreting and non-accreting stars described above can thus not be explained by differences in their rotation periods or their rotation-activity relations. Considering the relation between X-ray activity and Rossby numbers, we also find no significant correlations for either accretors or non-accretors. ### 8.3 X-ray activity and accretion rates/luminosities Robberto et al. (2004) recently determined accretion rates and accretion luminosities for a sample of 40 TTS in the Trapezium cluster from HST $`U`$\- and $`B`$-band photometry. As the computation of the accretion parameters from the UV excess is quite indirect and involves numerous assumptions, their values have considerable uncertainties; for some of their stars they find even negative values for the accretion luminosities. We therefore restrict us to those 30 stars for which they derived positive values for the accretion luminosity and note that 29 of these are detected as X-ray sources in COUP. For a considerable fraction of these objects the X-ray luminosities are comparable or even larger than the accretion luminosities; this is a strong argument against the assumption that that the X-ray emission in these TTS is directly created in the accretion process (e.g., comes from the accretion shock; see discussion below). Furthermore, we find a weak anti-correlation of the fractional X-ray luminosity with accretion rate (and also with accretion luminosity); although these correlations are not statistically significant, they agree to the above results based on the Ca line width classification and suggest that active accretion somehow lowers the X-ray activity levels. ### 8.4 Summary of the relations between X-ray activity and accretion We find that the TTS with inner circumstellar material as traced by near-infrared excess show slightly higher fractional X-ray luminosities than the TTS without near-infrared excess, but this difference is of only marginal statistical significance. Using the equivalent width of the Ca II line to discern between accreting and non-accreting stars, we find that the non-accretors show very well defined correlations between X-ray luminosity and bolometric luminosity or stellar mass. The accreting stars, on the other hand, produce much poorer correlations between $`L_\mathrm{X}`$ and $`L_{\mathrm{bol}}`$ or stellar mass and much more scatter. Furthermore, the mean fractional X-ray luminosities of the non-accreting TTS are well consistent with those of rapidly rotating MS stars, while the accreting TTS show about 3 times lower levels of X-ray activity. Finally, we have shown that X-ray activity appears to be anti-correlated with mass accretion rate. In conclusion, the X-ray activity of the non-accreting TTS is consistent with that of rapidly rotating MS stars, while in accreting TTS the X-ray emission is somehow suppressed. ## 9 Implications on the origin of TTS X-ray emission In this section we summarize what implications we can derive from our X-ray data on the origin of the observed X-ray emission from the TTS. We will consider the following questions: Where is the X-ray emitting plasma located? What is the reason for the lower X-ray activity of accreting stars in comparison to non-accretors? What is the ultimate origin of the magnetic activity and what kind of dynamo may work in the TTS? ### 9.1 Location of the X-ray emitting structures #### 9.1.1 X-ray emission from accretion shocks? According to the magnetospheric accretion scenario, accreted material crashes onto the stellar surface with velocities of up to several 100 km/sec, what should cause hot ($`10^6`$ K) shocks in which strong optical and UV excess emission and perhaps also soft X-ray emission is produced (e.g., Lamzin et al., 1996; Calvet & Gullbring, 1998). The expected characteristics of X-ray emission from accretion shocks would be a very soft spectrum (due to the low plasma temperature in the shock of at most a few MK), and perhaps simultaneous brightness variations at optical/UV wavelengths and in the X-ray band. Although earlier studies failed to find evidence for this scenario (e.g., Gullbring et al., 1997), more recent high-resolution X-ray spectroscopy of some TTS (e.g. TW Hya, XZ Tau and BP Tau, see Kastner et al., 2002; Stelzer & Schmitt, 2004; Favata et al., 2003; Schmitt et al., 2005) yielded plasma temperatures and densities that have been interpreted as evidence for X-ray emission originating from a hot accretion shock. The COUP results provide no support for a scenario in which X-ray emission is dominated by accretion power. First, we note that many of the accreting TTS show X-ray luminosities considerably larger than their total accretion luminosities. Although we have to be somewhat cautious because the X-ray and accretion rate measurements were not simultaneous and accretion is thought to be strongly time variable, it appears extremely unlikely that the bulk of the observed X-ray emission from the TTS could originate from accretion processes. Second, the COUP spectra of nearly all TTS show much higher plasma temperatures (typically a $`10`$ MK cool component and $`20`$ MK hot component) than the $`13`$ MK expected from shocks for the typical accretion infall velocities. Only for five of the TTS in our optical sample the X-ray spectral fits yielded plasma components with temperatures below 3 MK. For none of these stars accretion rates are known, only two of them show Ca II emission, and only one displays infrared excess. Thus, we cannot determine whether in any of these objects the X-ray emission may be related to accretion shocks. We just note that the low plasma temperatures alone provide no evidence for an accretion shock origin of the X-rays, since similarly low (or even lower) plasma temperatures are found for MS stars and the Sun, i.e. for stars that are certainly not accreting. Furthermore, the COUP lightcurves show thousands of high-amplitude flares whose temporal structure closely resembles solar-type magnetic flares, and is very unlikely to be reproduced by thermal accretion shocks. Indeed, Stassun et al. (2005) have compared these COUP lightcurves with simultaneous optical lightcurves and find very little evidence to suggest that X-ray variability is correlated with accretion-related processes. Of course, these arguments do not exclude the possibility that accretion shocks may contribute some fraction of the X-ray emission in TTS. It is critical to note that $`Chandra`$’s ACIS-I instrument is not very sensitive to the cooler plasma expected from these accretion shocks, and much of this emission may be attenuated by line-of-sight interstellar material. We note that evidence for a scenario of mixed X-ray emission in accreting YSOs has recently come from an X-ray high spectral resolution observation of BP Tau (Schmitt et al., 2005), where hot plasma (too hot to be shock-produced and thus likely magnetically confined and heated in some form of coronal structure) coexists with cool (1–3 MK) plasma for which unusually high densities were inferred, which may be well explained by accretion shock models<sup>10</sup><sup>10</sup>10The cool plasma in BP Tau shows a very low value of the ratio between the intensity of the forbidden and intercombination lines ($`R=f/i`$) for the O vii triplet, formed at temperatures of 1–3 MK. Low $`R`$ values (also observed in TW Hya, and in no coronal source) can either imply very high densities, or the presence of an intense UV field, as indeed expected within the accretion spot. Too high densities would however be difficult to explain given the pressure structure of an accretion shock, (see Drake, 2005), in which the plasma at $`n_e10^{13}`$ cm<sup>-3</sup> (the density implied by the $`R`$ value observed in TW Hya) is buried (following e.g. the shock structure and emission models by Calvet & Gullbring 1998) under a column density of typically $`10^{22}\mathrm{cm}^3`$, that should absorb the soft X-ray emission from the shock zone nearly completely and thermalize the high-energy radiation within or close to the shock.. We conclude that, although accretion shock emission must be present, plays an important role in optical or ultraviolet emission of CTT stars (e.g. Lamzin et al., 1996), and may contribute some fraction of the largely-unobserved soft X-ray emission of TTS, it is not responsible for the bulk of the X-ray emission seen in the COUP data. #### 9.1.2 X-ray emission from magnetic star-disk interactions? Another possibility for a non-solar like origin of the X-ray radiation may be plasma trapped in magnetic fields that connect the star with its surrounding accretion disk. The dipolar stellar magnetic field lines anchored to the inner part of the accretion disk should be twisted around because of the differential rotation between the star of the disk. Theoretical work suggests that the twisted field lines periodically reconnect, and the released magnetic energy heats plasma trapped in the field lines to very hot, X-ray emitting temperatures (Hayashi et al. 1996; Montmerle et al. 2000; Isobe et al., 2003; Romanova et al., 2004). The analysis of the $`30`$ largest flares in COUP data by Favata et al. (2005) suggests that very long magnetic structures (up to a few times $`10^{12}`$ cm) are present in some of the most active stars in our sample. Such large structures (tens of times the size of the star) may indicate a magnetic link between these stars and their disks. However, we note that very large loop lengths were derived for only a few of these flares; for the majority of the analyzed flares much smaller loop lengths were found. Furthermore, our results in the previous sections show that, in general, the X-ray luminosity is strongly linked to stellar parameters like bolometric luminosity and mass, but does not strongly depend on the presence or absence of circumstellar disks as traced by near-infrared excess emission. It is reasonable to conclude that although X-ray emission from magnetic star-disk interactions may be seen during some of the most intense flares, the bulk of the observed X-ray emission from ONC TTS probably originates from more compact structures with geometries resembling solar coronal fields. #### 9.1.3 X-ray emission from solar-like coronal structures? Our data are generally consistent with the assumption that the observed X-ray emission originates from, in principle solar-like, coronal structures. The X-ray luminosities and plasma temperatures derived for the ONC TTS are comparable to those of the most active MS stars, and can thus be understood by assuming that stellar coronae in general are composed of X-ray emitting structures similar to those present in the solar corona (e.g. Drake et al. 2000; Peres et al. 2004). Although the high X-ray luminosities of active MS stars and TTS can not be reproduced by simply filling the available coronal volume with solar-like active regions, coronal structures with higher plasma density<sup>11</sup><sup>11</sup>11Remember that the emission measure, $`EM:=n^2dV`$, scales linearly with the plasma volume $`V`$, but with the square of the density $`n`$. can explain the high levels of stellar X-ray activity. Evidence for higher than solar plasma densities is found in high-resolution X-ray and EUV spectra for many active stars (e.g. Sanz-Forcada et al., 2003; Ness et al., 2004). Furthermore, once the stellar coronae get nearly completely filled with active regions, the magnetic interaction of the active regions should lead to increased flaring in the most active stars, boosting their X-ray luminosities even further (Peres et al. 2004). We also note that the temporal behavior of most flares seen in the COUP data is rather similar to what is seen in flares on the Sun or active MS stars (Wolk et al. 2005). Further evidence suggesting enhanced solar-like coronal activity as the source of the X-ray emission from active MS stars and TTS is summarized e.g., in Feigelson & Montmerle (1999) and Favata & Micela (2003). ### 9.2 The suppression of X-ray emission by accretion Our COUP data confirm previous results that accreting TTS show lower levels of X-ray activity than non-accretors (Stelzer & Neuhäuser 2001; Flaccomio et al. 2003; Stassun et al. 2004a). Here we discuss some possible explanations for this effect. Two previously suggested explanations can essentially be ruled out by our data. The first one is the suggestion that the systematically higher extinction of the accreting CTTS (due to absorption in the accretion disk) may be responsible for their weaker observed X-ray emission as compared to the WTTS (e.g. Stassun et al., 2004a). In our COUP data, individual extinction-corrected X-ray luminosities could be determined in a self-consistent fitting analysis of the individual X-ray spectra. The different levels of extinction in the accreting and non-accreting stars should not lead to errors in the extinction-corrected X-ray luminosities. We can also exclude the idea that accreting TTS are weaker X-ray emitter because their rotation is braked by magnetic disk locking, leading to weaker dynamo action and therefore less X-ray emission than in the non-braked WTTS. We have shown in §5.1 that neither the accretors nor the non-accretors show a relation between rotation and X-ray activity, and thus the difference in rotation rates cannot be the reason for the difference in the level of X-ray emission. #### 9.2.1 Accretion changes the magnetic field structure? Numerical simulations suggest that the pressure of the accreting material can distort the large-scale stellar magnetic field considerably (e.g. Miller & Stone, 1997; Romanova et al., 2004) and the magnetospheric transfer of material to the star can give rise to instabilities of the magnetic fields around the inner disk edge and cause reconnection events. The presence of accreting material also leads to higher densities in (parts of) the magnetosphere. In contrast to the WTTS, which probably have loops with moderate plasma densities, some fraction of the magnetic field lines in CTTS would be mass loaded and therefore have much $`(100\times )`$ higher densities. If now a magnetic reconnection event liberates a certain amount of energy, this can heat the plasma in the low-density loops of WTTS to X-ray emitting temperatures ($`10`$ MK), while the denser plasma in the mass loaded loops of the CTTS would be only heated to much lower temperatures, and remain too cool to emit X-rays. This effect may cause the lower X-ray luminosities of the CTTS as compared to WTTS. However, a more quantitative assessment of this model is difficult. According to magnetospheric accretion models, the fraction of the stellar surface that is covered by accretion funnels should be at most a few percent (e.g., Calvet & Gullbring, 1998; Muzerolle et al., 2001). This may be a too small fraction to explain the reduction of the X-ray luminosity by a factor of $`2`$. On the other hand, we note that the estimates of the area of accretion funnels are uncertain, and other factors like global changes in the topology of the magnetic field may play a (more?) important role. It therefore seems possible that magnetospheric accretion streams are somehow related to the different X-ray activity levels of accreting and non-accreting stars. It is also interesting to note that the analysis of the largest flares in the COUP data by Favata et al. (2005) seems to indicate that intense, active accretion may inhibit magnetic heating of the accreting plasma, while in stars which are not actively accreting the long magnetic structures may acquire a “coronal” character. #### 9.2.2 Accretion changes the stellar structure? The accretion process may alter the internal stellar structure and the differential rotation patterns, and thereby influence the magnetic field generation process. For example, Siess, Forestini, & Bertout (1999) found in their stellar evolution calculations that accretion reduces the efficiency of convection. This theoretical result agrees with another finding based on comparison of orbital masses of PMS stars with evolution models by Stassun et al. (2004b), who found that TTS seem to have lower convection efficiencies than MS stars. The reduced convection efficiency may lead to weaker dynamo action. Another effect may be that the magnetospheric star-disk coupling affects the differential rotation pattern at the stellar surface. If the coupling happens closer to the stellar equator than to the poles, the magnetospheric braking effect thought to be at work in CTTS may reduce the amount of differential rotation, and this also may decrease the efficiency of the dynamo. Finally, the magnetic star-disk interaction may also have an effect on the coupling between inner and outer layers within the star, and thereby affect the level of magnetic activity (see Barnes, 2003a, b). ### 9.3 Implications for pre-MS magnetic dynamos The MS activity-rotation relation is well-established in stars (e.g., Pallavicini et al., 1981; Pizzolato et al., 2003) and is usually interpreted in terms of the $`\alpha \mathrm{\Omega }`$-type dynamo that is thought to work in the Sun. A simplified description of the rather complicated processes by which this dynamo generates surface magnetic fields can be summarized as follows (for details see, e.g., Schrijver & Zwaan, 2000; Ossendrijver, 2003): The strong differential rotation in the tachocline, a region near the bottom of the convection zone in which the rotation rate changes from being almost uniform in the radiative interior to being latitude dependent in the convection zone, generates strong toroidal magnetic fields. While most of the toroidal magnetic flux is stored and further amplified in the tachocline, instabilities expel individual flux tubes, which then rise through the convection zone, driven by magnetic buoyancy, until they emerge at the surface as active regions. The power of the dynamo (i.e. the magnetic energy created by the dynamo per unit time) depends only indirectly on the rotation rate. The $`\alpha \mathrm{\Omega }`$ dynamo is principally dependent on the radial gradient of the angular velocity in the tachocline and the characteristic scale length of convection at the base of the convection zone. The empirical relationship between X-ray luminosity and rotation rate in MS stars does therefore not mean that the power of the dynamo scales with rotation rate, but rather that faster rotating stars have stronger velocity shear in the thin tachoclinal layer between the radiative core and the outer convective zone. The presence of an $`\alpha \mathrm{\Omega }`$-type dynamo at the bottom of the convection zone does not prevent other dynamo processes from also operating in a star. In the Sun, small scale turbulent dynamo action (e.g. Durney et al., 1993) is taking place throughout the convection zone and is thought to be responsible for the small-scale intra-network fields. Recent results (Bueno et al., 2004) suggest that the total magnetic flux generated by the small-scale turbulent dynamo action is much larger than previously assumed. This means that two conceptually distinct magnetic dynamos are simultaneously operating in the contemporary Sun. In the case of the Sun, the coronal activity is most likely dominated by the tachoclinal dynamo action. Most of the ONC TTS, however, are thought to be fully convective, or nearly fully convective, so the tachoclinal layer is either buried very deeply, or does not exist at all. It is therefore reasonable to assume that in these (nearly) fully convective TTS, a convective dynamo is the main source of the magnetic activity. An interesting possible alternative explanation may be that the conventional wisdom, i.e. that TTS are fully convective, is not correct. We note that several studies have shown that accretion can significantly change the stellar structure. For example, Prialnik & Livio (1985) found that even for moderate accretion rates the stars are no longer fully convective. More recently, Wuchterl & Tscharnuter (2003) found that accreting PMS stars are not fully convective; their model of a solar mass star at 1 Myr has a radiative core and a convective envelope, resembling the present Sun rather than a fully convective object. These results open the possibility for a solar-type tachoclinal dynamo to work in the TTS. Our results on the relation between X-ray activity and Rossby number are not inconsistent with that possibility. We provide in this study various empirical relationships of the X-ray luminosities with, e.g., stellar mass, or bolometric luminosity, which should also be relevant to the dynamos operating in TTS. A purely empirical explanation of these correlations is given by the existence of upper and lower limits to the X-ray activity levels, in analogy to results for MS stars. The upper limit is caused by the saturation level of magnetic activity around $`\mathrm{log}\left(L_\mathrm{X}/L_{\mathrm{bol}}\right)3`$ (e.g., Pizzolato et al., 2003). A lower limit is suggested by studies of nearby field stars, that led to the conclusion that all cool dwarf stars are surrounded by X-ray emitting coronae with a minimum X-ray surface flux of about $`1\times 10^4\mathrm{erg}/\mathrm{cm}^2/\mathrm{sec}`$ (Schmitt, 1997; Schmitt & Liefke, 2004); for early M-type TTS in the ONC this surface flux corresponds to $`\mathrm{log}\left(L_\mathrm{X}/L_{\mathrm{bol}}\right)6`$. The restriction of $`\mathrm{log}\left(L_\mathrm{X}/L_{\mathrm{bol}}\right)`$ to the range between about $`6`$ and $`3`$ leads to correlation between X-ray luminosity and bolometric luminosity; the correlation between X-ray luminosity and stellar mass can then be explained by the dependence of bolometric luminosity on stellar mass. An alternative explanation for the correlations can be based on the finding that the fractional X-ray luminosities increase with stellar mass (§4.2). This is consistent with the results of Pizzolato et al. (2003), who showed that in low-mass MS stars the saturation level in $`L_\mathrm{X}/L_{\mathrm{bol}}`$ increases with stellar mass. These results suggest a similar origin of X-ray activity in the TTS and MS stars, and thus provide support for the standard $`\alpha \mathrm{\Omega }`$ solar-type-dynamo model for TTS X-ray emission. ## 10 Summary The main results from our study of the X-ray properties of the TTS in the ONC can be summarized as follows: In the COUP data we detect X-ray emission from essentially every late-type (F to M) ONC star. There is no indication for the existence of an “X-ray quiet” population of stars with suppressed magnetic activity. We find that the X-ray luminosities of the TTS are correlated to bolometric luminosities, stellar masses, and effective temperatures. The $`L_\mathrm{X}M`$ correlation for the TTS shows a slope similar to the corresponding correlation for MS stars, which is probably related to the association between mass and MS X-ray saturation levels. Together, these lines of evidence suggest that the $`L_\mathrm{X}M`$ relationship may be more physically fundamental than X-ray relationships to bolometric luminosity, surface area, or rotation. Our data indicate a correlation between X-ray activity and rotation period, apparently in strong contrast to the well established anti-correlation seen for MS stars. However, the efficacy of our analysis is limited since rotation periods are only known for about 40% of the TTS in our sample, and the missing stars (i.e. those with unknown rotation periods) probably introduce a bias. If we consider Rossby numbers, we find that all TTS are located in the saturated or super-saturated regime of the activity$``$Rossby number relation for MS stars. In principle, the TTS may thus follow the same relation between X-ray activity and Rossby number as MS stars, but the large scatter in $`L_\mathrm{X}/L_{\mathrm{bol}}`$ at any given Rossby number suggests that other factors are also involved in determining the level of X-ray activity. The enormous scatter we generally find in the correlations between X-ray activity and other stellar parameters is larger than what one would expect due to X-ray variability, uncertainties in the variables, and the effects of unresolved binaries. Therefore, this wide scatter must be related to intrinsic differences in the individual TTS, and we find here that the influence of accretion on the X-ray emission seems to play an important role. There is a remarkable contrast between the X-ray properties of accreting and non-accreting stars: Our data confirm previous results that accreting stars are less X-ray active than non-accreting stars (although a statistically significant difference is only found for stars in the $`0.20.5M_{}`$ mass range) and suggest an anti-correlation between fractional X-ray luminosity and accretion rate. The non-accreting TTS have the same median X-ray activity level as rapidly rotating MS stars and show good $`L_\mathrm{X}L_{\mathrm{bol}}`$ and $`L_\mathrm{X}M`$ correlations with a scatter as expected from the uncertainties, X-ray variability, and unresolved binaries. The accreting TTS, on the other hand, show about 3 times lower X-ray activity levels and produce much less well defined $`L_\mathrm{X}L_{\mathrm{bol}}`$ and $`L_\mathrm{X}M`$ correlations with much wider scatter. These findings imply that the apparent X-ray deficit of the whole TTS sample (i.e. the median fractional X-ray luminosity of $`\mathrm{log}\left(L_\mathrm{X}/L_{\mathrm{bol}}\right)3.5`$, which is below the saturation limit around $`\mathrm{log}\left(L_\mathrm{X}/L_{\mathrm{bol}}\right)3.0`$ typically found for rapidly rotating MS stars) is solely due to the reduced X-ray activity of the accreting TTS. We discuss several possible explanations for the suppression of X-ray emission in accreting stars. The effect may be related to changes of the coronal structure or the internal stellar structure induced by the accretion process. We favor the idea that magnetic reconnection can not heat the dense plasma in mass-loaded accreting field lines to X-ray temperatures. The geometry of X-ray producing magnetic fields is also still uncertain. Solar-type coronal loops are probably the dominant source of the observed X-ray emission. However, we note that the study of the most powerful flares seen in COUP stars (Favata et al. 2005) suggests that in some objects star-disk field lines extending $`>10\times R_{}`$ from the stellar surface may be involved. Accretion shocks at the stellar surface can not be responsible for the emission seen in COUP sources. Finally, the ultimate origin of the X-ray activity of the TTS is most likely either a turbulent dynamo working in the stellar convection zone, or, if theoretical suggestions that accreting TTS may not be fully convective are correct, a solar-like $`\alpha \mathrm{\Omega }`$ dynamo at the base of the convection zone. COUP is supported by $`Chandra`$ Guest Observer grant SAO GO3-4009A (E. Feigelson, PI). Further support was provided by the Chandra ACIS Team contract NAS8-38252. We would like to thank L.A. Hillenbrand for useful comments on the manuscript, J.H.M.M. Schmitt and C. Liefke for information on the NEXXUS database, H. Peter for enlightening discussion about the solar corona, M. Hünsch for information about the X-ray luminosities of subgiants, and H. Shang for discussions about accretion processes. YCK has been supported by Korean Research Foundation Grant KRF-2002-070-C00045. BS, EF, GM and SS acknowledge financial support from an italian MIUR PRIN program and an INAF program for the years 2002-2004. Facility: CXO(ACIS)
warning/0506/astro-ph0506668.html
ar5iv
text
# Statistics of Voids in the 2dF Galaxy Redshift Survey ## 1 Introduction There are many tests to constrain the models of structure formation, which range from the large-scale statistics such as correlation functions (Peebles, 1980; Davis & Peebles, 1983; Zehavi et al., 2002; Norberg et al., 2002) or the power spectrum (Peebles, 1980; Tegmark et al., 2004; Cole et al., 2005), to detailed studies of the physical properties of individual galaxy clusters and voids. Although dense environments (i.e., galaxy clusters and groups) have been extensively studied, the underdense regions like giant voids attract less attention. Yet, they are not less important and can provide important information on galaxy formation (Hoyle et al., 2005; Goldberg et al., 2005; Rojas et al., 2005; Croton et al., 2005) and can give independent constraints on cosmological models (Peebles, 2001; Plionis & Basilakos, 2002; Croton et al., 2004; Colberg et al., 2005; Solevi et al., 2005; Conroy et al., 2005). Large voids have woken up more and more interest since their first detections 25 years ago (Kirshner et al. 1981, Rood 1981). At present, thanks to the advent of larger redshift surveys like the 2dFRGS (Colless et al., 2001) and the SDSS (York et al., 2000), higher resolution of cosmological simulations and better analytical frameworks, we can extract accurate statistical information about voids. This information can be used in different ways but one of the most important is to test the models of structure formation. Voids can be studied in different ways. One of the classical methods is the Void Probability Function (VPF, White, 1979; Fry, 1986), which gives the probability that a randomly located sphere of a given radius contains no galaxies (see e.g. Einasto et al 1991, Croton et al 2004, Solevi et al 2005). The number density of voids with radius greater than $`R`$ is another useful statistic. This number density and also the void significance can be estimated analytically (see Patiri et al. 2004). So far, it has been done only using numerical simulations or mock catalogs (Colberg et al., 2005). Voids in the 2dFGRS were previously studied by Hoyle & Vogeley (2004) and Croton et al. (2004). Croton et al. (2004) have measured the VPF for volume-limited galaxy samples covering the absolute magnitude range $`M_{b_\mathrm{J}}5\mathrm{log}_{10}h=18`$ to $`22`$. Their work mainly focused on the study of the dependence of the VPF on the moments of galaxy clustering as a test to discriminate among different clustering models. They found that the VPF measured from the 2dFGRS is in excellent agreement with the paradigm of hierarchical scaling of the galaxy clustering. In addition, they showed that the negative binomial model gives a good approximation of the 2dF data over a wide range of scales. On the other hand, Hoyle & Vogeley (2004) have also calculated the VPF in the 2dFGRS obtaining similar results. They have obtained the VPF for the dark matter matter halos in $`\mathrm{\Lambda }`$CDM simulations and galaxy mock catalogs from semi-analytic models of galaxy formation to compare with the data. They have found that the results from the semi-analytic models that include feedback effects provide a VPF that agree with the VPF measured for the 2dFGRS and differ from that measured from the dark matter distribution. In spite of the fact that the notion of voids is not new, there is no standard definition of what is a void. “Voids” sometimes mean quite different objects. It all depends on used data and goals of the analysis. For example, to explain the patterns of the galaxy distribution in the Universe Van de Weygaert & Van Kampen (1993) and Sheth & Van de Weygaert (2004) define voids as irregular low-density regions in the density field. Colberg et al. (2005) use a similar definition to study void properties in a $`\mathrm{\Lambda }`$CDM universe. However, as these definitions are not based on point distributions, it may be difficult to deal with the galaxy samples provided by large scale redshift surveys. El-Ad H. & Piran (1997), Hoyle & Vogeley (2002) and Hoyle & Vogeley (2004) define voids as irregular regions of low number-density of galaxies, which may contain bright galaxies. Thus, by construction, voids are not empty even of very luminous and likely massive galaxies. By contrast, Gottlöber et al. (2003) define voids as spherical regions which do not have massive objects (halos in this case). Voids also can be defined in a statistical point of view as maximal spheres (Otto et al., 1986; Einasto, Einasto, & Gramann, 1989; Gottlöber et al., 2003; Patiri, Betancort-Rijo & Prada, 2004). In this paper we define voids as the maximal non-overlapping spheres empty of objects with mass (or luminosity) above a given value . For example, we could define voids as maximal spheres empty of Milky Way-size galaxies. While voids are empty of these galaxies, they could have fainter galaxies inside. See Figure 1 for a graphical representation of our definition. The first step to follow using voids as a test for large scale structure and galaxy formation models is to develop a robust algorithm to detect them, to calibrate it, and to obtain the statistical properties of voids for different catalogs both real and simulated ones. With this information and with the predictions made through the analytical formalism we may be able to contrast different structure formation models. To achieve this goal we develop two algorithms. They are conceptually different but are based on the same definition of void. We develop them as complementary tools. One algorithm is intended to search for all the voids in galaxy or dark matter halo samples and the other is developed to search for the rarest voids. Once we have the tools to detect the voids, we study statistical properties that will be used to test the structure formation models. These properties go from the void correlations to the redshift dependence of voids. In the present work, we apply the tools we mentioned above in order to study the statistic of voids in the 2dF Galaxy redshift Survey. One important point that could provide clues on the galaxy formation processes is the galaxy contents of voids. In this work, we present the first results on the distribution of faint galaxies in nearby voids. In section 2 we briefly describe how to detect voids. In section 3 we present the statistics of voids found in cosmological numerical simulations and compare the performance of our algorithms. In section 4 we present the voids that we have found in the 2dFRGS together with their statistical properties. In section 5 we develop a method to get from the redshift space coordinates, the real space ones. In section 6 we present the statistics of the galaxies inside nearby rare voids. finally, in section 7 conclusions and discussions are presented. In Appendix A we give details of our void finders presented in section 2. Here we also provide tests in order to check the performance of them. Throughout this work we adopt a $`\mathrm{\Lambda }`$CDM cosmology model with parameters $`\mathrm{\Omega }_m=0.3`$ and $`\mathrm{\Omega }_\mathrm{\Lambda }=0.7`$. ## 2 Void Detection Algorithms Once we have defined what is a void (as the maximal non-ovelapping spheres empty of galaxies brighter than a given magnitude), the next step is to develop an algorithm to be able to find them in galaxy or dark matter halo samples. The computation structure of the algorithm naturally will depend mainly on the void definition. Following the definition we have stated in the previous section, the algorithm should try to find the maximal sphere that the void can accommodate. There are many algorithms in the literature inspired by the different void definitions (see e.g. Einasto et al 1989, Kauffman & Fairall 1991, El Ad & Piran 1997, Aikio & Mahoenen 1998, Gottlöber et al. 2003, Colberg et al 2004). Aikio & Mahoenen 1998 and Colberg el al 2004 have developed similar algorithms. Following their definitions of voids, they generate a field smoothing the point distribution on a cubic grid and find the local minima; then, they put spheres around those minima filling the entire underdense region. The radius of these voids are the effective ones taken from the sphere that contain the same volume as the void. Gottlöber et al. 2003 have developed an algorithm based on the minimal spanning tree for halos selected by intrinsic properties, in this case with mass above a given value. Although they have searched for voids as local maximal spheres, they did not do voids statistics in their work. Another algorithm in literature is the so-called void finder developed by El Ad & Piran (1997) and its modifications done by Hoyle & Vogeley (2002). This algorithm search for arbitrarily shaped regions delimited by the so-called wall galaxies in order to get the maximum volume of the void. They classify the galaxy sample in wall galaxies and field or voids galaxies by mean of a criteria that depends on the galaxy distribution itself, i.e. they define a length parameter $`l_n`$ such that any galaxy that does not have $`n`$ neighbours within a sphere of radius $`l_n`$ is classify as field galaxy. Note that field galaxies could be, for example, bright galaxies. These field galaxies are removed before searching for voids. Once the classification is done, the algorithm search for the maximal spheres defined by the wall galaxies based on a cubic mesh. Once all the spheres are obtained, they define an overlapping parameter to discriminate if two spheres belong to the same void. Similarly to Aikio & Mahoenen (1998) and Colberg el al (2004), they finally fill the voids with spheres to get the maximum volume. Again, they define the effective radius of a void by means of a sphere that contain the same volume as the void. Note that, although this may be interesting for some studies (e.g. the shapes of voids), these kind of statistics degrade in some amount the information available in the actual galaxy distribution, so they might not be particularly powerful to conduct accurate statistical inferences. This is similar to what happens with the binned data: the best statistical test using binned data is never better and usually worst than the best test using the raw data. To identify voids, we designed two algorithms that are conceptually different but both are based in our definition of voids. The algorithms are complementary. They help us to investigate the variety of aspects present in the statistics of voids. One of the algorithms, which we call CELLS void finder, was designed to search for all the voids in a galaxy or halo sample based on a computational grid. This grid define the working resolution. To determine the void centers, the code computes the distances between each of the empty grid cells and all the galaxies or halos in the whole observational domain, keeping the minimum distance. Once we have the list with the minimum distances, we search for the local maxima which corresponds to the void centers. Obviously, the voids radius are those maximum distances. The other algorithm, which we denominate HB Void Finder is conceptually simple. This code searches for the maximal non-overlapping spheres with radius larger than a given value. First of all, we generate over the sample a large sample of random spheres of a given radius. After this, we check and keep the spheres that are empty of galaxies. We inflate these spheres until they reach the maximum radius. Finally, we eliminate the overlapping spheres keeping the maximal ones. This code is very accurate and computationally efficient to search the biggest voids in a galaxy sample. Detailed descriptions of the two algorithms are given in the Appendix. ## 3 Void statistics in simulations ### 3.1 Numerical simulations We perform a series of numerical simulations with the Adaptive Refinement Tree code (ART, Kravtsov et al. 1997) and the Tree-SPH code GADGET (Springel, Yoshida & White, 2001). Dark matter halos are identified in the simulation by the Bound Density Maxima algorithm (BDM, Klypin & Holtzman 1997; Klypin et al. 1999). In this work we detect and study voids in simulation boxes of $`80h^1\text{Mpc}`$, $`120h^1\text{Mpc}`$ and $`500h^1\text{Mpc}`$ size. The parameters of all the simulations are summarized in Table 1. With these boxes we have enough volume to study accurately the void statistics and compare the results obtained with our two void finders. ### 3.2 The statistics In Figure 2 we show the number density of voids as a function of their radius obtained with our void finders (squares for CELLS void finder and circles for the HB void finder). These results were obtained for voids defined by two halo masses ($`1\times 10^{12}\mathrm{M}_{_{\mathrm{}}}`$ (open symbols) and $`5\times 10^{12}\mathrm{M}_{_{\mathrm{}}}`$ (filled symbols)). In order to search for voids with the HB Void Finder, we have generated $`1\times 10^7`$, $`2\times 10^7`$, $`5\times 10^7`$ trial spheres for the $`80h^1\text{Mpc}`$, $`120h^1\text{Mpc}`$ and $`500h^1\text{Mpc}`$ boxes respectively. In the case of the Cell Void Finder we defined a working resolution of $`0.5h^1\text{Mpc}`$ for the $`80h^1\text{Mpc}`$ and $`120h^1\text{Mpc}`$ boxes and $`1.2h^1\text{Mpc}`$ for the $`500h^1\text{Mpc}`$ box. The obtained statistics shows that both algorithms gives very similar results. These results are indeed very useful to learn about the performance of the void finders. Both algorithms give exactly the same results for the largest voids (these are rare voids), being the HB Void Finder the fastest and more precise. However, as we go to more common voids the Cell Void Finder has the best performance. The small differences between both codes for smaller voids are due to the fact that we need more realizations of the trial spheres in the HB Void Finder. As it is expected, the voids are more numerous and larger the more massive the halos defining the void are. ## 4 Void statistics in the 2dF Galaxy Redshift Survey ### 4.1 The galaxy samples In the present work we use the 2dFGRS final data release (Colless et al. 2003) to obtain the void statistics in large galaxy redshift samples. The source galaxy catalogue of the 2dFRGS is taken from the APM galaxy catalogue (Maddox et al. 1990). The spectroscopic targets are galaxies with extinction corrected magnitudes brighter than $`b_J=19.45`$. The median depth of the survey is $`z0.11`$. The final data releases contains a total of 221,414 high quality redshifts. There are two large contiguous survey regions, one in the south Galactic pole (SGP) and another one towards the north Galactic pole (NGP). There are also a number of random fields which we have eliminated from our void search. Full details of the 2dFGRS can be found in Colless et al.(2001, 2003). In order to search for the voids we have selected from 2dFGRS two rectangular regions: the region in the SGP defined by $`34^{}40^{}<\delta <25^{}12^{}`$ and $`21^\mathrm{h}49^{}<\alpha <3^\mathrm{h}26^{}`$ and the region in the NGP defined by $`4^{}35^{}<\delta <2^{}17^{}`$ and $`9^\mathrm{h}33^{}<\alpha <14^\mathrm{h}54^{}`$; that is, $`690`$ and $`550`$ sq. degrees respectively. The 2dFGRS is magnitude-limited, i.e. the survey has been constructed by taking spectra of galaxies brighter than a fixed apparent magnitude of $`b_\mathrm{J}=19.45`$. However the survey is homogeneously complete up to 90% at $`b_\mathrm{J}`$=19.0 (see Norberg et al. 2002). A magnitude-limited galaxy survey is not uniform in space, since intrinsically faint galaxies have been observed only if they are relatively nearby, while at large distances only bright galaxies will be targeted. This non uniformity of the magnitude-limited survey must be taken into account in order to make our void analysis. There are mainly two ways to deal with this; one is to use the selection functions provided by the 2dFGRS and another more simple way which we have followed here is to build volume-limited samples. We construct four volume-limited samples, two for each survey region (SGP and NGP), one with depth $`D_{max}=406.15h^1\text{Mpc}`$ which corresponds to $`z=0.14`$. For this $`D_{max}`$ we have a limiting absolute magnitude $`M_{b_\mathrm{J}}^{lim}=19.32+5\mathrm{l}\mathrm{o}\mathrm{g}h`$ and the other with depth $`D_{max}=571.71h^1\text{Mpc}`$ corresponding to $`z=0.2`$ and a limiting absolute magnitude $`M_{b_\mathrm{J}}^{lim}=20.181+5\mathrm{l}\mathrm{o}\mathrm{g}h`$. All distances are comoving ones. In Table 2 we give the properties of the volume-limited samples. Now, we have guaranteed that any galaxy brighter than $`M_{b_\mathrm{J}}^{lim}`$ is observed in our volume. We have computed the absolute magnitudes from the apparent magnitudes assuming a $`\mathrm{\Lambda }`$CDM cosmology and applying the needed corrections to model the change in the galaxy magnitudes due to the redshifted $`b_\mathrm{J}`$ filter bandpass (k-correction) and to account for the galaxy evolution (e-correction). These corrections for each galaxy are given in Norberg et al.(2002). To test the spatial homogeneity of our SGP1,2 and NGP1,2 volume-limited samples we have computed the average of the cube of the radial distances, i.e. a modified version of the $`V_{max}`$ test (Rowan-Robinson 1968), which for a homogeneous sample must satisfy: $`<\left({\displaystyle \frac{r}{r_{max}}}\right)^3>={\displaystyle \frac{1}{2}}\pm {\displaystyle \frac{1}{\sqrt{12N^{}}}}`$ (1) $`N^{}={\displaystyle \frac{N}{(1+N<\xi (r)>)}}`$ where $`r`$ is the comoving distance to each galaxy, $`r_{max}`$ is the same as $`D_{max}`$, the maximum distance of the sample, $`N`$ is the number of galaxies (given in table 2) and $`<\xi (r)>`$ is the average value of the correlation function over all pair of galaxy positions within the sample, defined as: $$<\xi >=\frac{1}{V^2}\xi (|𝐫_\mathrm{𝟏}𝐫_\mathrm{𝟐}|)𝑑𝐫_\mathrm{𝟏}𝑑𝐫_\mathrm{𝟐}$$ (2) where $`V`$ is the volume of the sample. Assuming for $`\xi (r)`$ (Peebles 1980): $$\xi (r)=\left(\frac{r}{5.4h^1\text{Mpc}}\right)^{1.77}$$ (3) we found $`<\xi (r)>=6.7910\times 10^3`$. It must be noted that we have used eq.(3) for any value of $`r`$, while in fact, it is known that for $`r>>10h^1\text{Mpc}`$ $`\xi (r)`$ must be close to the fourier transformed of the linear power spectra. This would lead to smaller values of $`<\xi (r)>`$. However, this would not affect our analysis too much. With this value we find that for a homogeneous sample: $$\eta <\left(\frac{r}{r_{max}}\right)^3>=\frac{1}{2}\pm 0.024.$$ (4) The actual $`\eta `$ values for the SGP1 and NGP1 are $`\eta (SGP1)=0.524`$ and $`\eta (NGP1)=0.51`$. For the SGP2 and NGP2 to be homogeneous (assuming that $`<\xi (r)>`$ takes the same values as in the previous sample): $$\eta <\left(\frac{r}{r_{max}}\right)^3>=\frac{1}{2}\pm 0.0253,$$ (5) and the actual values are $`\eta (SGP2)=0.511`$ and $`\eta (NGP2)=0.475`$. Therefore, we can conclude that we do not detect inhomogeneities. ### 4.2 The void statistics As described in section 3.2, once we have constructed our samples, we have to define which objects will define the voids. In the SG1 and NGP1 volume limited samples we present in table 2 we search for voids defined by two types of galaxies. In the one hand, we search voids defined by galaxies brighter than $`M_{b_\mathrm{J}}=19.32+5\mathrm{l}\mathrm{o}\mathrm{g}h`$. Using these samples we also find voids defined by galaxies brighter than $`M_{b_\mathrm{J}}=20.5+5\mathrm{l}\mathrm{o}\mathrm{g}h`$. The number of galaxies brighter than $`M_{b_\mathrm{J}}=20.5+5\mathrm{l}\mathrm{o}\mathrm{g}h`$ is 2427 for the SGP1 sample and 2074 for the NGP1. In the SGP2 and NGP2 samples, we have searched for voids defined by galaxies brighter than $`M_{b_\mathrm{J}}=20.181+5\mathrm{l}\mathrm{o}\mathrm{g}h`$. In Figure 3 we show a plot with the voids we have detected in the NGP1 sample. In table 3 we summarize the results for the NGP1 and SGP1 samples. Note that, as we saw in the simulations, voids defined by bright galaxies are larger than the ones defined by fainter ones. In table 4 we present the statistics for the NGP2 and SGP2 samples. We have obtained these voids applying the HB algorithm. In the radius ranges that we present here both codes have a similar performance. We generate $`8\times 10^7`$ trial spheres to search for voids larger than $`7.5h^1\text{Mpc}`$ which are our main interest. We have also calculated the VPF which is the probability that a randomly located sphere of fixed radius contains no galaxies. Figure 4 presents our VPF, which we find for the SGP1,2 and NGP1,2 samples. We calculate the rms of the VPF in the following way. Assuming that voids are independent, we get from the central limit theorem that $$rms^2(P_0(r))=\frac{N(r)}{V^2}(<\mathrm{\Delta }v^2><\mathrm{\Delta }v>^2)$$ (6) where $`\mathrm{\Delta }v`$ is the volume where the center of an empty sphere of radius $`r`$ may be moved so that it remains empty. $`V`$ is the volume of the sample and $`N(r)`$ is the number of voids found with radius larger than $`r`$ in that volume. The parenthesis in the right hand side of eq.(6) divided by $`V^2`$ is the contribution of each individual void to the variance of the VPF. On the other hand we have from Betancort-Rijo (1992) that $$\frac{N(r)}{V}=\frac{P_0(r)}{<\mathrm{\Delta }v>}.$$ (7) Now we square this expression and substitute it in expression (6) to obtain: $$rms^2(P_0(r))=\frac{P_0^2(r)}{N(r)}(g(r)1)$$ (8) $$g(r)\frac{<\mathrm{\Delta }v^2>}{<\mathrm{\Delta }v>^2}$$ (9) So that, taking into account the correlations in the sample, the rms of the VPF is given by: $$rms(P_0(r))g(r)^{1/2}\frac{P_0(r)}{N(r)^{1/2}}(1w\overline{n}_v(R))^{1/2}$$ (10) where $`w\overline{n}_v(R)`$ is as defined in Eq.(21). In the rare voids limit, i.e. when $`P_0<<<1`$ it is shown that (Betancort-Rijo, 1992): $$g(r)9.20$$ (11) For voids with radius larger than $`12h^1\text{Mpc}`$ in the SGP1 and NGP1 samples and larger than $`17h^1\text{Mpc}`$ in the SGP2 and NGP2 samples, the asimptotic value $`g(r)^{1/2}=2.86`$ is a good approximation. For smaller values of $`r`$, $`g(r)^{1/2}`$ is somewhat smaller (no less than $`2`$). So in figure (4), where the asimptotic value of $`g(r)`$ have been used, the error may be slightly overstimated for small values of $`r`$. It must be noted, however, that the probability distribution of the fluctutations of the VPF around the mean is strongly non-Gaussian for small values of $`N(r)`$; most fluctuation being quite smaller than the $`rms`$ and a few of them being very large. So, to decide the compatibility between couples of measurements, this fact has to be taken into account. In the present case, however, this problem do not arize because all results are within the error bars) Our results are in good agreement with the previous calculations of the VPF (Hoyle & Vogeley 2004, Croton et al. 2004). In Croton et al.(2004) the VPF was obtained for the whole 2dFGRS while Hoyle & Vogeley (2004) have studied the VPF for both strips (NGP and SGP). In this last result, they found that the VPF for both strips are not compatible for large spheres (see figure 7 in Hoyle & Vogeley 2004). This is be due to the fact that they underestimated the error bars. We have shown above that using a more accurate expression for the rms of the VPF we obtain that both strips are compatible. ### 4.3 Voids spatial distribution If we neglect the redshift dependence of the power spectra, the spatial distribution of the voids found in a statistically homogeneous sample must be statistically homogeneous itself, but conditioned to the fact that the maximal sphere must lay within the sample. This fact implies that the center of the maximal spheres can not occupy all the volume of the sample. The available volume, $`V(R)`$, for the maximal spheres of radius $`R`$ is: $`V(R)`$ $`=`$ $`{\displaystyle _{\frac{R}{\mathrm{sin}(\frac{\mathrm{\Delta }\delta }{2})}}^{r_{max}R}}{\displaystyle _{\delta _0}^{\delta _0+\mathrm{\Delta }\delta }}{\displaystyle _{\alpha _0+B(r,R,\delta )}^{\alpha _0+\mathrm{\Delta }\alpha B(r,R,\delta )}}r^2\mathrm{cos}\delta 𝑑\alpha 𝑑\delta 𝑑r`$ (12) $`B(r,R,\delta )`$ $`=`$ $`\mathrm{sin}^1\left({\displaystyle \frac{\mathrm{sin}^1(\frac{R}{r})}{\mathrm{cos}\delta }}\right)`$ Eq.(12) may be considerably simplified by carrying out the integrals over $`\alpha `$ and $`\delta `$ keeping $`B`$ fixed (using its value at $`\delta _0+\mathrm{\Delta }\delta /2`$). This is a very good approximation since $`\mathrm{\Delta }\delta <<1`$. So, we could write: $$V(R)_{\frac{R}{\mathrm{sin}(\frac{\mathrm{\Delta }\delta }{2})}}^{r_{max}R}P(r,R)r^2𝑑r$$ (13) $`P(r,R)`$ $``$ $`[\mathrm{sin}(\delta _0+\mathrm{\Delta }\delta \mathrm{sin}^1({\displaystyle \frac{R}{r}}))\mathrm{sin}(\delta _0)]\times `$ (14) $`\times `$ $`\left[\mathrm{\Delta }\alpha 2\mathrm{sin}^1\left({\displaystyle \frac{\mathrm{sin}^1(\frac{R}{r})}{\mathrm{cos}(\delta +\frac{\mathrm{\Delta }\delta }{2})}}\right)\right]`$ $`\mathrm{\Delta }\delta `$,$`\mathrm{\Delta }\alpha `$ are, respectively, the widths of the strips in declination and right ascension, while $`\delta _0`$ is its southernmost declination. Although expression (14) for V(R) is not exact, its percentual error is completely negligible (less than $`10^5`$). Using $`P(r)`$ we may obtain the mean of the cube of the distance (from the observer), $`r`$, to the center of the maximal sphere with radius larger than $`R^{}`$: $`\eta (R^{})`$ $``$ $`<\left({\displaystyle \frac{r}{r_{max}}}\right)^3>_{RR^{}}=`$ $`=`$ $`{\displaystyle \frac{1}{V(\overline{R})}}{\displaystyle _{\frac{\overline{R}}{\mathrm{sin}(\frac{\mathrm{\Delta }\delta }{2})}}^{r_{max}\overline{R}}}P(r,\overline{R})\left({\displaystyle \frac{r}{r_{max}}}\right)^3r^2𝑑r`$ where $`\overline{R}`$ is the mean radius of all maximal spheres larger than $`R^{}`$. In a similar manner we may obtain the variance for the estimate of $`\eta (R/h^1\text{Mpc})`$ from the void list we have obtained in the 2dFGRS. The values found in the SGP1,2 and NGP1,2 for voids defined by galaxies brighter than $`19.32+5\mathrm{l}\mathrm{o}\mathrm{g}h`$ are listed in table 5). From the analysis of these results we may conclude that the voids are essentially compatible with homogeneity. There is, however, a slight, although statistically significant, trend to values lower than homogeneous that we will discuss in subsection (3.5). Now, we will consider the nearest neighbour statistics which together with the void one point statistic we have studied above is sufficient to grant the validity of the statistical analysis in the following subsections. The maximal spheres are chosen so that they do not overlap, in order that they very rarely corresponds to the same connected underdensity. So, by construction, for the maximal spheres with radius larger than $`R`$ the two point correlation function of their centers is $`1`$ at least to a distance of $`2R`$. The main question is how is the correlation at larger distances. Note that if more than one maximal sphere were associated with the same underlying connected underdensity the correlation would be positive for distances somewhat larger than $`2R`$. However, as there is typically only one maximal sphere per underdense region and in the standard scenario of structure formation these regions are essentially uncorrelated (for the relevant $`R`$ values)(Colberg et al. 2004), we do not expect to find correlations between the maximal spheres. To test whether voids are correlated we compare the actual nearest neighbour statistics with the theoretical predictions corresponding to centers which are uncorrelated for distances larger than $`2\overline{R}`$ and completely anticorrelated for smaller distances. For maximal spheres with radius $`R`$ the theoretical predictions for the mean and quadratic mean distance to the nearest neighbour ($`<D_{oo}>`$ and $`<D_{oo}^2>`$ respectively) is given by: $$<D_{oo}>=\left(\frac{4\pi \overline{n}_v(R)}{3}\right)^{1/3}_{w_0}^{\mathrm{}}e^{(ww_0)}w^{1/3}𝑑w$$ (16) $`<D_{oo}^2>`$ $`=`$ $`\left({\displaystyle \frac{4\pi \overline{n}_v(R)}{3}}\right)^{2/3}{\displaystyle _{w_0}^{\mathrm{}}}e^{(ww_0)}w^{2/3}𝑑w`$ (17) $`w_0`$ $``$ $`{\displaystyle \frac{4\pi }{3}}(2\overline{R})^3\overline{n}_v(R)`$ where $`\overline{n}_v(R)`$ is the mean number density of the maximal spheres larger than $`R`$ and $`\overline{R}`$ is the mean radius. The sampling error of $`<D_{oo}>`$ when estimated using $`N`$ centers is: $$rms(<D_{oo}>)=\frac{\sqrt{5}(<D_{oo}^2><D_{oo}>^2)^{1/2})}{(3N2)^{1/2}}$$ (18) with these expressions we may compute the mean and the variance of $`q`$: $$q\left(\frac{R}{h^1\text{Mpc}}\right)(\overline{n}_v(R))^{1/3}E(<D_{oo}>)$$ (19) where $`E(<D_{oo}>)`$ is the estimate of $`<D_{oo}>`$ using the void sample. For typical voids we find for $`q`$: $`q_{SGP1}(7.5)`$ $`=`$ $`0.978(0.958\pm 0.006)`$ $`q_{NGP1}(7.5)`$ $`=`$ $`1.029(0.958\pm 0.008)`$ $`q_{SGP2}(13.0)`$ $`=`$ $`0.966(0.879\pm 0.01)`$ $`q_{NGP2}(13.0)`$ $`=`$ $`1.1(0.879\pm 0.013)`$ where $`R=7.5h^1\text{Mpc}`$ ($`\overline{R}=9.58h^1\text{Mpc}`$) correspond to SGP1, NGP1 and $`R=13.0h^1\text{Mpc}`$ ($`\overline{R}=15.08h^1\text{Mpc}`$) to SGP2, NGP2. The values in the parenthesis correspond to $`1\sigma `$ predictions obtained using Eqns.(17) and (18). It is apparent that the estimated values are slightly shifted upwards. The bias is large in terms of the $`rms`$’s but we can not conclude from this that the centers are anticorrelated since the predictions do not account for border effect, which in this cases is small for the SGP samples but larger than the $`rms`$. On the other hand, for rare voids, i.e. $`R=13.0h^1\text{Mpc}`$ for the SGP1, NGP1 samples ($`\overline{R}_{SGP1}=14.14h^1\text{Mpc}`$,$`\overline{R}_{NGP1}=13.42h^1\text{Mpc}`$) and $`R=18.0h^1\text{Mpc}`$ for the SGP2, NGP2 samples ($`\overline{R}_{SGP2}=19.5h^1\text{Mpc}`$,$`\overline{R}_{NGP2}=18.45h^1\text{Mpc}`$) we find: $`q_{SGP1}(13.0)`$ $`=`$ $`0.959(0.716\pm 0.004)`$ $`q_{NGP1}(13.0)`$ $`=`$ $`1.398(0.716\pm 0.006)`$ $`q_{SGP2}(18.0)`$ $`=`$ $`0.971(0.724\pm 0.046)`$ $`q_{NGP2}(18.0)`$ $`=`$ $`1.071(0.724\pm 0.070)`$ The upward bias with respect to the values for the uncorrelated centers (without accounting for border effects) is now large even in the SGP. To asses the border effect we have simulated a random set of spheres with $`R=19.5h^1\text{Mpc}`$ (the mean radius of the maximal spheres with radius $`18.5h^1\text{Mpc}`$ in the SGP2 sample) with the condition that they lay entirely within the catalogue and do not overlap, then we obtain the nearest neighbour statistics. The average over several realizations of the mean nearest distance is: $$\overline{n}_v(18.0)<\overline{D}_{oo}>=0.88\pm (0.05\pm ϵ)$$ (20) which is quite larger than the value predicted without border effect (0.724). This is not enough, however, to account for the measured value (0.971). The same result is found for $`z_{max}=0.14`$ since the border effect in this case must be very close to that in the previous case. From these results we might conclude at least at the 99% confidence level (being rather conservative about the value of $`ϵ`$) that the center of the maximal spheres show some anticorrelation. We prefer, however, to conclude simply that those centers are basically uncorrelated, which is the necessary result for the subsequent statistical inferences, and leave open the question of the possible existence of a weak anticorrelation, which shall be treated in more detail in a future work. Incidentally, it must be noted that a slight anticorrelation of the centers arise naturally in the standard scenario, being compatible with the results of Colberg et al. 2004. The reason is that the conexed underdense regions are somewhat larger than the maximal spheres they contain. ### 4.4 Compatibility between SGP and NGP samples Here we compare the void statistics found in our samples in order to asses their compatibility. To this end we must estimate the mean number density of voids, $`\overline{n}_v(R)`$, in each sample and obtain their sampling error. To estimate $`\overline{n}_v(R)`$ we simply divide the number of voids,$`N(R)`$, with radius larger than $`R`$, by the mean available volume $`V(\overline{R})`$. To estimate the sampling error we must take into account that, although the center of the maximal spheres are basically uncorrelated for $`r>2\overline{R}`$, they are totally anticorrelated for $`r<2\overline{R}`$ (we use a model in which all spheres have radius $`\overline{R}`$). In this case the number of spheres in a given volume do not follow a Poissonian distribution, but a negative binomial one (Betancort-Rijo 1999). For this distribution the sampling error of the estimate of $`\overline{n}_v(R)`$ could be obtained by means of the following expression: $$rms(\overline{n}_v(R))=\frac{\overline{n}_v(R)}{(N(R))^{1/2}}(1w\overline{n}_v(R))^{1/2}$$ (21) $`w{\displaystyle \frac{32\pi }{3}}\overline{R}^3`$ $`[`$ $`1+2.73\overline{n}_v(R){\displaystyle \frac{32\pi }{3}}\overline{R}^3`$ $`\times `$ $`(1{\displaystyle \frac{3}{4}}({\displaystyle \frac{\overline{n}_v(R)^{1/3}}{2\overline{R}}})+{\displaystyle \frac{1}{8}}({\displaystyle \frac{\overline{n}_v(R)^{1/3}}{2\overline{R}}})^2)]^1`$ For SGP1 we find: $`\overline{n}_v(7.5)`$ $`=`$ $`9.364\pm 0.164\times 10^5`$ $`\overline{n}_v(10.0)`$ $`=`$ $`3.98\pm 0.11\times 10^5`$ $`\overline{n}_v(13.0)`$ $`=`$ $`7.26\pm 1.70\times 10^6`$ For NGP1 we find: $`\overline{n}_v(7.5)`$ $`=`$ $`9.597\pm 0.214\times 10^5`$ $`\overline{n}_v(10.0)`$ $`=`$ $`4.27\pm 0.15\times 10^5`$ $`\overline{n}_v(13.0)`$ $`=`$ $`6.56\pm 2.10\times 10^6`$ The corresponding $`\overline{R}`$’s are: $`\overline{R}(7.5)=9.58`$ and $`\overline{R}(10.0)=11.35`$ (the mean over NGP and SGP), $`\overline{R}(13.0)=14.14`$ in the SGP1 and $`\overline{R}(13.0)=13.41`$ in the NGP1. It is apparent that the three pair of values are compatible. It is true that for the SGP1 the galaxy density is roughly a 10% smaller than for the NGP1 and that the theoretical expression (Patiri, Betancort-Rijo & Prada, 2004) predict about a 20% enhancement of the density of voids with $`R13h^1\text{Mpc}`$ in the SGP1 over that in the NGP1, and smaller differences for more common voids, but this differences do not show up above the sampling errors. Similar results were found for the SGP2: $`\overline{n}_v(13.0)`$ $`=`$ $`1.685\times 10^5`$ $`\overline{n}_v(15.0)`$ $`=`$ $`9.49\times 10^6`$ $`\overline{n}_v(18.0)`$ $`=`$ $`2.978\pm 0.58\times 10^6`$ and NGP2, $`\overline{n}_v(13.0)`$ $`=`$ $`2\times 10^5`$ $`\overline{n}_v(15.0)`$ $`=`$ $`1.202\times 10^5`$ $`\overline{n}_v(18.0)`$ $`=`$ $`2.22\pm 0.69\times 10^6`$ Although the implications of this results for the large scale structure formation models is left to for a future work, it is interesting to note that they are generally in good agreement with the predictions of the standard model. ### 4.5 Redshift dependence of void number densities We have checked that the galaxies in our samples are, within the sampling errors, homogeneously distributed over the sample volumes. In this situation, the only possible explanation of an statistically significant departure of the void distribution from the homogeneity must be the redshift dependence of the properties of the large scale structure. Using a theoretical framework (Patiri, Betancort-Rijo & Prada, 2004) to compute the number density of voids in the distribution of dark matter halos, we found that the predicted number density of voids larger than $`13.0h^1\text{Mpc}`$ for $`M_{b_\mathrm{J}}^{lim}=19.32+5\mathrm{l}\mathrm{o}\mathrm{g}h`$ decreases by 28% from z=0 to z=0.1 and a further 28% when going to z=0.2. For $`M_{b_\mathrm{J}}^{lim}=20.181+5\mathrm{l}\mathrm{o}\mathrm{g}h`$ and $`R18h^1\text{Mpc}`$ the void density decreases by 26% from z=0 to z=0.1 and another 28% from z=0.1 to z=0.2. Where, in order to link light with dark matter halos, we assumed that there is one galaxy in each dark halo. These numbers suggest the possibility of a measurable effect. To this end we conduct first a test where we divide every sample into two bins (a ’near’ region and a ’far’ region) both with the same volume available for the voids considered. Then we compare the total number of voids larger than the chosen radius in the near regions with the total number in the far regions. If the first number is significantly larger than the latter this should be interpreted as evidence for a redshift dependence. The problem now is that the effect we are trying to measure is strong only for the rare voids which have poorer statistics. We reach a compromise between both trends. We have chosen voids with radius ($`R_{min}`$) larger than $`13h^1\text{Mpc}`$ for the $`z_{max}=0.14`$ sample and $`18h^1\text{Mpc}`$ for the $`z_{max}=0.2`$ sample. In order to divide a sample into two subsamples with the same volume available for the centers of the voids, we first obtain the mean radius ($`\overline{R}`$) of the voids larger than $`R_{min}`$, and, using Eq.(12) with $`R=\overline{R}`$, we obtain the availabe volume of the whole sample (which has already been done in section 4.3). Then, we use again (12) with $`R=\overline{R}`$ but replacing $`r_{max}`$ with $`r_{lim}+\overline{R}`$ and search for the value of $`r_{lim}`$ giving for $`V`$ half the value corresponding to the whole sample. If voids were uniformly distributed their centers would lie above and bellow $`r_{lim}`$ with equal probability. We have explicitly checked this fact with numerical simulations. For the sample SGP1, we have chosen $`R_{min}=13h^1\text{Mpc}`$, so $`\overline{R}=14.14h^1\text{Mpc}`$ which imply $`r_{lim}=371.7h^1\text{Mpc}`$. 9 voids were found with their centers bellow this distance and 2 above. Proceeding in a similar way with the other samples, we found 3 voids closer than $`349.1h^1\text{Mpc}`$ and 3 voids further for the NGP1. For the SGP2 we found 5 voids closer than $`475.0h^1\text{Mpc}`$ and 8 further, while for the NGP2 we found 4 voids closer than $`490.2h^1\text{Mpc}`$ and 1 further. So, in total we have 21 voids in near regions and 14 in the far regions. With these numbers we can infer at least at the 88% confidence level the presence of a redshift dependence. However, this is only a marginal evidence. To make more patent this evidence we use a more efficient test. This test uses the $`\eta `$ values obtained in 4.3 which are shown in table 5. We use the $`\eta `$ values correpsonding to voids larger than $`13h^1\text{Mpc}`$ in the NGP1 and SGP1 samples, and voids larger than $`18h^1\text{Mpc}`$ in the NGP2 and SGP2 samples. From each of those four $`\eta `$ values we substract its expected value when uniformity is assumed (first number in the parenthesis in table 5) and divide the result by the corresponding *r*ms (second number in the parenthesis). Each of these quantities follow a Gaussian distribution with zero mean and variance 1 under the uniformity hipothesis. So, the sum of the four quantities must follow a Gaussian with zero mean and variance 4 and the probability that these sum be smaller than the actually found result is: $$1\frac{1}{2}erfc\left(\frac{_{i=1}^4\frac{\eta _i\overline{\eta _i}}{rms(\eta _i)}}{2\sqrt{2}}\right)=0.005$$ (22) So, we may infer at the 99.5% confidence level the existence of non-uniformity on the void distribution, that, given the uniformity of the galaxy distribution can only be explained by the growth of density fluctuations as redshift decreases. Note that the $`rms(\eta )`$ given in table 5 do not account for the anticorrelation between the voids centers (when $`r<2R`$). The last factor in expression (21) approximately accounts for this fact. If we were to use it, all the $`rms(\eta )`$ would be smaller and, consequently, we could reject the uniformity hypothesis at a larger confidence level. However, we prefer to give the conservative value quoted above since it has not any uncertaninty. ## 5 Redshift Space Distortions Redshift galaxy maps like the 2dFGRS or SDSS are distorted by the peculiar velocities of galaxies along the line of sight. This effect, which produce deformations such as the “finger of God”, is due to the fact that the distance to the survey galaxies is obtained by means of the Hubble distance which is obtained with the total velocity (the Hubble flow plus the peculiar velocity). We present in this section a method to correct this effect for the galaxies around voids. We first suppose that if in the distorted space we have a void of radius $`R`$, we will have the same void of radius $`R`$ in the real space. This is a good approximation because the distortion around a void cause an elongation along the line of sight of the maximal sphere without changing the “transversal” radius, i.e., although the distortion is not volume conservative, the maximal radius will be approximately the same in both spaces. With this assumption, we have in the position of the void in the distorted space a mean underdensity, $`\overline{\delta }_0`$, within the maximal sphere with radius $`R`$ (Patiri, Betancort-Rijo & Prada, 2004): $$\overline{\delta }_0=_1^{\mathrm{}}\delta _0P(\delta _0/R)𝑑\delta _0$$ (23) where $`P(\delta _0/R)`$ is the probability distribution for $`\delta _0`$ within a maximal sphere of radius $`R`$. We associate with this void a matter distribution given by the mean profile, $`\delta (r/\overline{\delta _0},R)`$. In Patiri, Betancort-Rijo & Prada (2004) we derive this profiles. So, it has a peculiar velocity profile $`V(\delta (r),r)`$ given by the spherical collapse model (Betancort-Rijo et al. 2005): $$\frac{V(\delta (r),r)}{H}=\frac{0.51}{3}\frac{r}{1+\delta }\delta _l(\delta )\left(\frac{d\delta _l(\delta )}{d\delta }\right)^1$$ (24) where $`H`$ is the Hubble constant, $`\delta _l(\delta )`$ is given in Sheth & Tormen (2002) and $`\delta `$ is the mean profile (note that for simplicity we do not show the dependence on $`r`$, $`\delta _0`$ and $`R`$). We have derived an analytical approximation for $`\delta (r/\overline{\delta _0},R)`$ which gives good results for $`r1.5R`$: $$\delta (r/\overline{\delta _0},R)\overline{\delta _0}+(0.1645+0.085\delta _0)\left(\frac{r}{1.4R}\right)^7$$ (25) So that, we correct the galaxy “measured” distance $`r(z)`$: $$r_{real}r(z)\frac{(\stackrel{}{x}\stackrel{}{x}_c).\stackrel{}{x}}{|(\stackrel{}{x}\stackrel{}{x}_c)||\stackrel{}{x}|}\frac{V(\delta (r/\overline{\delta _0},R),r)}{H}$$ (26) where $`\stackrel{}{x}`$ is the vector to the galaxy and $`\stackrel{}{x}_c`$ is the vector to the center of the void. We apply this correction to galaxies with distance to the center of the void $`1.5R`$. Once we have the corrected catalogue, we search for voids over this new catalogue. In order to test our technique, we have applied it to a simulated catalogue, the Millennium Run galaxy catalogue (Springel et al 2005). The public available catalogue<sup>1</sup><sup>1</sup>1It can be downloaded from http://www.mpa-garching.mpg.de/galform/agnpaper/ contains a total of about 9 million galaxies in the simulation box of $`500h^1\text{Mpc}`$. For each galaxy, it is available the position and velocity, the total and bulge galaxy magnitudes in 5 bands (ugriz SDSS bands), the total and bulge stellar mass, cold, hot and ejected gas mass, the black hole mass and the star formation rate. The dark matter halos in the simulation were populated using semi-analytic models of galaxy formation (see Croton et al. 2005 for full details). We have constructed from the full simulated box a smaller one of $`250h^1\text{Mpc}`$. With this box we have enough volume to study large (and rare) voids. Also, as we have available in the original catalogue the coordinates in real space, we have constructed another box of $`250h^1\text{Mpc}`$ with the same galaxies but with the coordinates in redshift space. We have applied our HB void finder to search for voids larger than $`13.0h^1\text{Mpc}`$ in both distorted and real catalogues. With the list of voids in the distorted catalogue, we have applied the correction to those galaxies that lie within $`1.4`$ radius of the voids. After we have obtained the corrected galaxy catalogue, we run the void finder over this catalogue. In table (6) we show the statistics of voids in the 3 different catalogs: the distorted, the real and the corrected. From the study of the simulated catalogues we learn that, although the correction may be larger than $`3h^1\text{Mpc}`$ (for $`R16h^1\text{Mpc}`$) for galaxies close to the line of sight to the center of the voids, the correction for the radius of the maximal sphere is, as expected, much smaller. Even so, the difference is not negligible for sufficiently rare voids. We find that the 20 largest voids in the simulated catalogues are on average slightly smaller ($`5`$%) in the corrected catalogue. From the void statistics found in the corrected SGP1 and NGP1 samples we find that for the corrected SGP1 the ten largest voids are on average $`0.83h^1\text{Mpc}`$ smaller than for the uncorrected one and that the number of voids larger than $`12h^1\text{Mpc}`$ in both samples is 25 for the uncorrected case and 14 for the corrected one. So, since the strongest constraint on the models comes from relatively rare voids, it seems likely that the corrected catalogues must be used in order to be able to make accurate inferences. ## 6 Galaxies in Nearby Voids In this section we study the galaxy content of the nearby rarest voids in our 2dFGRS samples. To this end, we have selected the voids from our SGP1 volume-limited sample described in Section 5.1 (see Table 3). The voids are defined by galaxies brighter than $`19.32+5\mathrm{log}h`$. We have searched for faint galaxies down to $`M_{b_\mathrm{J}}^{\mathrm{lim}}=18.3+5\mathrm{log}h`$ inside nine voids with radius larger than $`13h^1\text{Mpc}`$ in a volume limited sample up to z=0.095. These voids are uncommon (with mean radius of $`14.0h^1\text{Mpc}`$) due to the fact that a randomly placed sphere with radius equal to the mean size of these voids should contain about $`50`$ galaxies brighter than $`19.32+5\mathrm{log}h`$. In total we find 130 faint galaxies inside these 9 rare voids, i.e. on average 14 galaxies per void. We have estimated the number density contrast of the galaxies located inside these voids by $$\delta _{gal}=\frac{\overline{n}_{void}\overline{n}}{\overline{n}},$$ (27) where $`\overline{n}_{void}`$ is the number density of galaxies inside the void and $`\overline{n}`$ is the number density of galaxies in the field. For the galaxies inside the voids we obtain $`\delta _{gal}=0.87`$. In figure 5 we display a central $`6h^1\text{Mpc}`$ thick slice for three of these nine voids. We can see that, despite the fact that these voids are highly empty, the faint galaxies populating them show interesting structures like filaments. Nevertheless, most galaxies are placed close to the borders of the voids being their centers much emptier. Note that these galaxy patterns are similar to those found in voids in high resolution numerical simulations by Gottöber et al. 2003. In figure 6 we show the mean number density of faint galaxies in our rare voids as a function of distance to the void center (normalized to the void radius). We have also studied the galaxies inside common voids, i.e. voids whose underdensities are not too big. We have selected from the SGP1 sample the voids defined by galaxies brighter than $`20.5+5\mathrm{log}h`$. We found five voids larger than $`15.0h^1\text{Mpc}`$ up to z=0.095. For these radius, a randomly chosen sphere should contain only 3 galaxies brighter than $`20.5+5\mathrm{log}h`$. In total there are 666 galaxies fainter than $`20.5+5\mathrm{log}h`$ down to $`M_{b_\mathrm{J}}^{\mathrm{lim}}=18.3+5\mathrm{log}h`$. These voids contain on average 10 times more galaxies than the rare voids discussed above, being the number density contrast of galaxies $`\delta _{gal}=0.54`$. In figure 7 we show an example of a $`8h^1\text{Mpc}`$ thick slice of a $`18.25h^1\text{Mpc}`$ void in this sample. The galaxies inside this void almost fully fill the void (open circles). This is not surprising due to the fact that the density contrast of these voids is not too big. ## 7 Discussion & Conclusions We have developed two robust and accurate algorithms to detect non-overlapping maximal spheres in halo or galaxy samples. We have applied them to several numerical simulations boxes in order to study the performance of the algorithms. The results found were very satisfactory. We have then studied the void statistics in the 2dFGRS. We have detected $`350`$ voids with radius larger than $`7.5h^1\text{Mpc}`$ defined by galaxies with $`M_{b_\mathrm{J}}<19.32+5\mathrm{l}\mathrm{o}\mathrm{g}h`$ and $`70`$ voids with radius larger than $`13.0h^1\text{Mpc}`$ defined by galaxies with $`M_{b_\mathrm{J}}<20.5+5\mathrm{l}\mathrm{o}\mathrm{g}h`$ in the volume limited samples up to $`z_{max}=0.14`$. We have also obtained the void statistics for the volume limited samples up to $`z_{max}=0.2`$. For this case, we have detected $`170`$ voids with radius larger than $`13h^1\text{Mpc}`$ defined by galaxies brighter than $`M_{b_\mathrm{J}}<20.181+5\mathrm{l}\mathrm{o}\mathrm{g}h`$. The number density of voids larger than $`R`$ found in the SGP and NGP samples are in good agreement with each other for all values of $`R`$. We have obtained the VPF for both strips finding results in good agreement with previous ones (Croton et al., 2004; Hoyle & Vogeley, 2004). We have shown, using an appropriate expression for the VPF sampling errors, that the results found in both strips are statistically compatible. From the results of several statistical tests we conclude that except for the anticorrelation implicit in the fact that the maximal sphere are chosen so that they do not overlap, they are essentially uncorrelated. There is, however, some evidence for a weak additional anticorrelation, which may be easely explained within the standard scenario of structure formation. We conclude at least at 99.5% confidence level that there is a dependence of the void number density on redshift. We do this by means of a modified version of the $`V_{max}`$ test which reveals a small trend towards small z values. We have also obtain preliminary results on the galaxy contents of nearby voids found in volume-limited galaxy samples in the 2dFGRS. For the nine nearby voids up to $`z<0.095`$ in our sample, we have found inside them on average only 15 galaxies fainter than $`19.32+5\mathrm{log}h`$ (the magnitude of the galaxies which define the voids). These voids are rather empty compare to those defined by brighter galaxies. The galaxies within the voids are not randomly distributed: they are clustered forming well define filamentary structures as that observed in the large-scale structure of the galaxy distribution of the Universe. Moreover, this is the same pattern that show the dark matter halo distribution found inside voids in high resolution N-body simulations. The halos inside voids are distributed in a way that resemble a miniature of the Universe (see Figure 2 in Gottlöber et al. (2003)). We have also obtained the number density profile for the galaxies inside the voids. The number density of faint galaxies fall almost a factor 7 from the border of the voids to the inner half. Moreover, the number density of galaxies close of the borders of the voids are still too low compared with the field (almost a factor 5). Gottlöber et al. (2003) and Patiri, Betancort-Rijo & Prada (2004) have obtained the number density profile for halos in voids. From the comparison of these results we can see that even though the number density of galaxies and halos follow similar trends, the galaxy profile is steeper than halos (they fall just a factor 2). In a future work, we will use the framework developed in Patiri, Betancort-Rijo & Prada (2004) along with the results given here for the number density of voids and their redshift dependence in order to constrain the value of $`\sigma _8`$. Furthermore, using a halo occupation model (e.g. Berlind et al. 2003) along the lines described in Patiri, Betancort-Rijo & Prada (2004) we shall establish constraints in the relationship between halos and galaxies from void statistics and the statistics of the galaxies inside the voids. In particular, we hope to be able to determine whether or not the conditional luminosity function depends on environment. In the other hand, the study of the physical properties of the galaxies inside voids could imply constraints on the galaxy formation processes. So, we will analyze the physical properties like colors, metalicities, star formation rates, etc. of galaxies in rare voids that are available from the biggest galaxy surveys (2dFGRS and SDSS) in order to test, for example, the galaxy luminosity function in voids and the density-morphology correlations. ## Acknowlegments We thank the referee for the comments and suggestions which improved the previous version of the paper. S.G.P. would like to thank the hospitality of the Department of Astronomy of the New Mexico State University where part of this work was done. We also thank support from grant PNAYA 2005-07789. This work was also partially supported by the Acciones Integradas Hispano-Alemanas. A.K. acknowledges support of NSF and NASA grants to NMSU. Computer simulations have been done at CESGA (Spain), NIC Jülich, LRZ Munich and NASA. The Millennium Run simulation used in this paper was carried out by the Virgo Supercomputing Consortium at the Computing Centre of the Max-Planck Society in Garching. ## Appendix A Void finder algorithms ### A.1 Algorithm 1: Cell Void Finder The first step is to select the sample of galaxies or dark matter halos in our redshift survey or cosmological simulation that will define the voids. We will select galaxies with luminosity greater than a given value $`L`$ or halos with virial mass greater than $`M`$. Once the galaxy or halo sample has been selected our code generates a cubic mesh where the cell size determines the working resolution. Afterwards, the objects are assigned to cells. So we have three types of cells: the cells that contain galaxies or halos (filled cells), the cells that are empty but are inside the observational domain (observed empty cells) and those cells which are also empty but they are located outside the observational domain or in a not observed region inside the observational domain (a ’hole’) (not observed empty cells). The code searches for the voids among the observed empty cells inside the observational domain. In the case of cosmological simulations the dark matter halo samples are in 3D boxes and generally all the cells are inside the observational domain. However, in the case of galaxy samples in redshift surveys which have irregular geometry, some of the cells are located outside the observational domain. We can easily determine which cell has been observed and which has not by using the survey masks. Once the cell classification procedure have been completed the code is then ready to search for the voids, finding their centers and radii. In principle, each observed empty cell could be a potential void center, but it is easy to realize that the observed empty cells which are located close to filled cells that contain galaxies or halos will not be the center of a void. So, in order to save computational time, we mark these neighbours cells and they will not be taken into account at the time of searching for the void centers. This is an iterative process, i.e. once we have marked the observed empty cells closer to the filled cells we can go to the next level and mark the observed empty cells that are neighbours of already marked empty cells. Note that we will stop this iteration depending on the working resolution (see Figure A1). To determine the void centers, the code computes the distances between each of the unmarked observed empty cell and all the galaxies or halos in the whole observational domain and we retain the minimum distance. Once we have the list with the minimum distances, we search for the local maxima which corresponds to the void centers. Obviously, the voids radius are those maximum distances. Finally, the code removes the overlapping maximal spheres, keeping the biggest one, i.e. if the distance between two maximal spheres is less than the sum of their radius, then the voids overlap and we remove the smaller one. The main advantage of the Cell Void Finder algorithm is that in only one run we get all the voids in the sample. However, its main disadvantage is that it consumes quite a lot of memory, which scale with the resolution that we require. There is a similar memory problem when we have a large number of galaxies or halos in the sample. If we are only interested in the biggest voids (rare voids) this algorithm is not the best strategy. Some studies are focused in these kind of voids, so, here we have developed a complementary algorithm which is more efficient in this respect. ### A.2 Algorithm 2: HB Void Finder Here we give the details of our second algorithm whose main task is the detection of rare voids and we will use it as a complement of the CELLS algorithm. The HB Void Finder is conceptually simple. It searches for the non-overlapping maximal spheres with radius larger than a given value. As we mentioned above this code is designed for statistical studies focused on the biggest voids. In these cases the code is very accurate and computationally efficient as compared with our Cell Void Finder. The first step in the algorithm is to generate a sample of random trial spheres with a fixed radius $`R`$. These spheres are generated directly over the observational domain with the condition that the entire sphere lies inside the observational domain. We check which spheres contains no objects and keep them. In the next step we find for each trial empty sphere the four nearest object and we expand the sphere to contain them in its surface. These new spheres are potentially maximal (see subsection A4). Note that the new expanded spheres could in principle contain objects. If this is the case, those spheres are removed. The potential maximal spheres will be actual maximal spheres if they do not overlap, which is decided by means of the same criteria as for the Cell Void Finder (i.e. if the distance between two spheres is less than the sum of their radius they overlap, so we keep the largest one) and if the four nearest objects are not located in the same hemisphere (see Figure A2). We put in both algorithms the additional constrain that the maximal spheres have to be entirely inside the observational domain. Note that, we need to do many realizations of the trial spheres in order to get the maximal spheres. Typically, four realizations are needed for each void in order to reach its maximum radius, when the maximal sphere is only slightly larger than $`R`$ and an increasing number as the maximal sphere is larger with respect to $`R`$. ### A.3 Performance test We construct random samples of objects in order to test the code performance and check the results of both algorithms. We generate two samples, one with 1,000 points and another one with 10,000 points. Both samples are in a box of $`100\mathrm{Mpc}`$. We give in Table A1 the void statistics computed from both algorithms. Note that the agreement between codes is excellent. We obtain the void statistics using $`10^7`$ trial spheres with the HB Void Finder. The resolution of the Cell Void Finder is $`0.5\mathrm{Mpc}`$. The cpu time of the codes mainly depends on the number of particles, the number and radius of trial spheres in the case of the HB algorithm and on the number of cells (i.e. resolution) and the levels of neighbours cell marking for the Cell Void Finder. For example, in the case of the sample with 1,000 random particles and for the voids with radius larger than $`10\mathrm{Mpc}`$, the HB Void Finder takes $`1`$ hour for $`10^7`$ trials, while the Cell Void Finder takes $`3`$ hours with a resolution of $`0.5\mathrm{Mpc}`$. Notice that the voids with radius larger than $`10\mathrm{Mpc}`$ are common in this box, so the running time differences are not so big between both algorithms. However, if we search for voids with radius larger than $`12\mathrm{Mpc}`$, the HB Void Finder takes only 20 minutes, while the Cell Void Finder last the same $`3`$ hours. These tests were done in a Pentium IV processor (3.06 GHz clock and 2GB RAM) and in a Itanium-2 processor (1.5 GHz clock, 2GB RAM) giving both similar performances. ### A.4 How to grow the trial spheres To determine the sphere passing through the four nearest objects to an empty trial sphere we proceed as follow: We first take the two nearest objects (whose coordinates we denote as $`\stackrel{}{x}_1`$ and $`\stackrel{}{x}_2`$) and calculate the middle point $`(\stackrel{}{x}_1+\stackrel{}{x}_2)/2`$. From this point, we move along a vector in the plane containing the three nearest objects and perpendicular to $`\stackrel{}{x}_2\stackrel{}{x}_1`$ until we reach the point, $`q`$, where the distances to the third nearest object ($`\stackrel{}{x}_3`$) is the same as that to object $`1`$, then the distance between object 2 and $`\stackrel{}{q}`$ is also the same that the previous two (See figure A3). Then we need to solve: $$|\stackrel{}{x}_1\stackrel{}{q}(w_0)|=|\stackrel{}{x}_3\stackrel{}{q}(w_0)|$$ (28) where $$\stackrel{}{q}(w)=\frac{\stackrel{}{x}_1+\stackrel{}{x}_2}{2}+w\stackrel{}{j}$$ (29) and $$\stackrel{}{j}=\frac{\stackrel{}{e}_{13}(\stackrel{}{e}_{13}.\stackrel{}{e}_{12})\stackrel{}{e}_{12}}{|\stackrel{}{e}_{13}(\stackrel{}{e}_{13}.\stackrel{}{e}_{12})\stackrel{}{e}_{12}|}$$ (30) where $$\stackrel{}{e}_{12}\frac{\stackrel{}{x}_1\stackrel{}{x}_2}{|\stackrel{}{x}_1\stackrel{}{x}_2|};\stackrel{}{e}_{13}\frac{\stackrel{}{x}_1\stackrel{}{x}_3}{|\stackrel{}{x}_1\stackrel{}{x}_3|}$$ (31) with $`w_0(2R_0,2R_0)`$, $`R_0`$ is the radius of the trial sphere. Now, we repeat the same procedure described above but taking into account the fourth object, i.e. we move from $`\stackrel{}{q}(w_0)`$ perpendicularly to the plane of the figure A1 until we reach the point $`\stackrel{}{P}(t)`$ where the distance between $`\stackrel{}{x}_4`$ and $`\stackrel{}{P}(t)`$ is the same as the distance from $`\stackrel{}{x}_1`$ to $`\stackrel{}{P}(t)`$. So, $$\stackrel{}{P}(t)=\stackrel{}{q}(w_0)+t\stackrel{}{n}$$ (32) where $$\stackrel{}{n}=\stackrel{}{e}_{12}\stackrel{}{e}_{13}$$ (33) Solving $$|\stackrel{}{x}_1\stackrel{}{P}(t)|=|\stackrel{}{x}_4\stackrel{}{P}(t)|$$ (34) with $`t(2R_0,2R_0)`$, we finally obtain the coordinates of the center of the maximal sphere, $`\stackrel{}{P}(t)`$, and its radius $`R`$, which is simply given by $`|\stackrel{}{x}_1\stackrel{}{P}(t)|`$.
warning/0506/cond-mat0506730.html
ar5iv
text
# First-order transitions for very nonlinear sigma models. ## 1 Introduction One of the main predictions of the Renormalisation Group (RG) theory is what is called “universality”. This means that in great generality the nature of the phase transition between high-temperature and low-temperature phases and the corresponding critical exponents depend only on dimension, symmetry and the range of the interaction. Here range means short-range (finite-range or sufficiently fast decaying with distance) or long-range (slowly decaying at large distances). The classical Landau mean-field theory similarly predicts that the nature of the spontaneously broken symmetry determines the order of the transition. Although in many cases such RG predictions have been confirmed, there are some examples where, somewhat unexpectedly, first-order instead of the predicted second-order (or absence of any) transitions were observed, see e.g. . In some cases, such as the nearest-neighbour $`q`$-state Potts models, one might think that it is the nature of the broken (permutation) symmetry which governs whether there is a first-order (at high $`q`$) or a second-order (at low $`q`$) transition. But the generalization of this statement is hard to make. Indeed, as Onsager already knew, there seems to be no general method to predict whether a transition is first-order or second-order. For example, it was shown in (which extended the related work of ) that a 3-state Potts model in dimension two, with an interaction of finite but large range, has a first-order transition, while for the nearest-neighbour model the transition is of second order. In this contribution, we review results from where we exhibited a different class of models. Though they possess global rotation symmetries, they undergo first-order transitions, whereas the universality predictions of the RG suggest second-order transitions. More precisely, for short-range ferromagnetic, $`d`$-dimensional, rotation-invariant $`n`$-vector models standard lore (which as we here show can be violated) predicts the following (“universal”) behaviour: * If $`d=2`$, there is a unique Gibbs measure at any positive temperature. For $`n=2`$ (classical XY spins – or the “nonlinear sigma model” in field theoretical language) there is nevertheless an infinite-order transition between a low-temperature Kosterlitz-Thouless phase with slowly decaying correlations and a high-temperature phase with exponential correlation decay. For higher $`n`$ there is no phase transition. * If $`d=3`$ or higher, there is a second-order transition between a magnetized low-temperature phase and a high-temperature phase. If $`d=3`$, one has $`n`$-dependent critical exponents, in higher dimensions one obtains mean-field exponents. Below we present a rather wide class of models in which this standard lore is violated. ## 2 Notation and some background For general background on the theory of Gibbs measures we refer to . We will consider spin models defined on the lattice $`^d`$ with spins taking values on the $`n`$-dimensional unit sphere. We will use small Greek letters $`\sigma ,\eta ,\mathrm{}`$ to denote spin configurations in finite or infinite boxes. The nearest-neighbour Hamiltonians in a box $`\mathrm{\Lambda }`$, for which we take a $`d`$-dimensional torus, will be given by $$H^\mathrm{\Lambda }(\sigma )=\underset{i,j\mathrm{\Lambda }}{}U(\sigma _i\sigma _j)+\underset{i\mathrm{\Lambda }}{}h\sigma _i.$$ (1) Associated to these Hamiltonians $`H^\mathrm{\Lambda }(\sigma )`$ are Gibbs measures $$\mu ^\mathrm{\Lambda }(d\sigma )=\frac{1}{Z^\mathrm{\Lambda }}\mathrm{exp}[H^\mathrm{\Lambda }(\sigma )]\mu _0^\mathrm{\Lambda }(d\sigma )$$ (2) Here $`\mu _0^\mathrm{\Lambda }(d\sigma )`$ denotes the rotation-invariant product measure. The choice of the function $`U`$ of the inner product between the spins at neighbouring sites will determine our model. We will study only the case of nearest-neighbour interactions. The reason is that our method is based on the use of certain correlation inequalities, called chessboard estimates, see below. These inequalities hold once the measures $`\mu ^\mathrm{\Lambda }`$ have the *Reflection Positivity (RP)* property, see again below, which RP holds for the n.n. interactions, see . The choice $$U(x)=x$$ (3) provides the standard classical XY and Heisenberg models. Equivalently, as a function $`W`$ of the difference angle $`\theta `$ between neighbouring spins, this means choosing $$W(\theta )=U(cos(\theta ))=cos(\theta )$$ (4) ## 3 Results, and some remarks on proofs. In this section we describe our results on the nonstandard $`n`$-vector models. We start with the case of zero field $`h=0`$ and $`d=2,n=2`$, so we have classical $`XY`$ spins in two dimensions. This seems to be the first case which was considered in the literature as an example of the phenomenon we display. However, the arguments in that paper – which we here prove to be correct – were later contested . The original choice of was $$W(\theta )=(\frac{1}{2}(1+cos\theta ))^p$$ (5) with $`p`$ large enough. A simpler but essentially similar model was introduced in : $$W(\theta )=\{\begin{array}{cc}1& \text{ if }\left|\theta \right|<\epsilon ,\\ 0& \text{ otherwise,}\end{array}$$ (6) where the parameter $`\epsilon `$ is small enough. Both these potentials have the form of a deep (depth $`=1`$) and narrow (width $`\frac{1}{\sqrt{p}},`$ cf width $`=2\epsilon `$) well, compared to the standard, rather shallow-well, cosine shape. The second model is a square well (or a top hat) potential, in which the distinction between being in or out of the well is unambiguous, in the first model there is a slight arbitrariness, and one has to make a choice to fix it. First we notice that by the Mermin-Wagner theorem , all Gibbs measures are rotation-invariant, so that the spontaneous magnetisation is necessarily zero. This does not prevent, however, the presence of multiple Gibbs measures (as was known already from the model of , where a discrete, chiral, symmetry was shown to be broken). Our first result is about the square-well model: Theorem *For $`\epsilon `$ small enough, there is a transition temperature where two Gibbs measures, an “ordered” one and a “disordered” one, coexist. In the ordered state most bonds are ordered, in the sense that the two spins at its ends have a difference angle smaller than $`2\epsilon `$ (they are in the well), in the disordered state the opposite is true.* Remarks about the proof: The proof is a fairly straightforward application of the Reflection Positivity, chessboard estimates method, which was developed by Dyson, Fröhlich, Israel, Lieb and Simon . For the benefit of the reader we recall these concepts. Let $`R_L`$ be a reflection of our torus $`\mathrm{\Lambda }`$ in some plane $`L`$ passing through its sites. (To have such a symmetry plane, $`\mathrm{\Lambda }`$ has to be of even size.) Then $`L`$ cuts $`\mathrm{\Lambda }`$ into two halves, $`\mathrm{\Lambda }_+`$ and $`\mathrm{\Lambda }_{},`$ so that $`R_L\left(\mathrm{\Lambda }_\pm \right)=\mathrm{\Lambda }_{}.`$ Let $`\mathrm{\Lambda }_0=\mathrm{\Lambda }L;`$ in the case $`d=2`$ this intersection is a pair of meridians of $`\mathrm{\Lambda }.`$ Let us consider the conditional distribution of the measure $`\mu ^\mathrm{\Lambda }`$ under condition that the restriction $`\sigma _{\mathrm{\Lambda }_0}\sigma |_{\mathrm{\Lambda }_0}`$ is fixed. Then it is easy to see that for every value of $`\sigma _{\mathrm{\Lambda }_0}`$ the conditional measure $`\mu ^\mathrm{\Lambda }(|\sigma _{\mathrm{\Lambda }_0})`$ splits into a product of two identical measures, $`\mu ^{\mathrm{\Lambda }_\pm ,\sigma _{\mathrm{\Lambda }_0}}`$, living on corresponding halves, with $`R_L\left(\mu ^{\mathrm{\Lambda }_\pm ,\sigma _{\mathrm{\Lambda }_0}}\right)=\mu ^{\mathrm{\Lambda }_{},\sigma _{\mathrm{\Lambda }_0}}.`$ From that it follows immediately that for every function $`C\left(\sigma _\mathrm{\Lambda }\right)=C\left(\sigma _{\mathrm{\Lambda }_+}\right),`$ depending only on the “left” variables $`\sigma _{\mathrm{\Lambda }_+}`$, we have $$CR_LC𝑑\mu ^\mathrm{\Lambda }0.$$ (7) The Reflection Positivity property is precisely the validity of this inequality. The details can be found in , Theorem 17.21. Note that one can choose the symmetry plane $`L`$ arbitrary, so in fact we have many such inequalities, and one corollary of this set of inequalities is the following chessboard estimate. To describe the simplest example of such an estimate let us consider a random variable $`D\left(\sigma _\mathrm{\Lambda }\right),`$ which depends on just one spin value, $`\sigma _0,`$ where $`0\mathrm{\Lambda }`$ is the origin. Then for its expected value $`D\left(\sigma _0\right)_{\mu ^\mathrm{\Lambda }}D\left(\sigma _0\right)𝑑\mu ^\mathrm{\Lambda }\left(\sigma _\mathrm{\Lambda }\right)`$ we have $$D\left(\sigma _0\right)_{\mu ^\mathrm{\Lambda }}\left[\underset{\begin{array}{c}x\mathrm{\Lambda },\\ x\text{ is even}\end{array}}{}D\left(\sigma _x\right)_{\mu ^\mathrm{\Lambda }}\right]^{\frac{4}{\left|\mathrm{\Lambda }\right|}}.$$ (8) Here $`D\left(\sigma _x\right)`$ is the same function $`D`$, computed at value $`\sigma _x,`$ and we take a product over all sites $`x`$ with both coordinates even. Note that if the interaction $`U`$ is identically zero, then the measure $`\mu ^\mathrm{\Lambda }`$ is just the product measure $`\mu _0^\mathrm{\Lambda }(d\sigma ),`$ and the last inequality becomes an equality. Applying this kind of chessboard estimate to various observables it is fairly straightforward to show that at low temperatures most bonds are ordered and that at high temperatures most bonds are disordered. The main step then left is to prove that uniformly for all temperatures in a temperature interval, including both high and low temperatures, the probability for two arbitrary bonds to be different (one ordered, one disordered) is small. Indeed, the only way these two properties can hold simultaneously is the existence of an intermediate temperature at which both the ordered and the disordered phase coexist. We will explain now how an estimate of the probability that a certain bond $`b_1`$ is ordered, while another one, $`b_2,`$ is disordered, can be obtained. If such an event happens, then there exists a contour $`\gamma ,`$ separating $`b_1`$ and $`b_2,`$ formed by sites which have a pair of orthogonal bonds, one of them being ordered and another disordered. For example, one obtains such a contour by taking the appropriate component of the boundary of the set of ordered bonds, containing the bond $`b_1.`$ Therefore it is enough to obtain a Peierls-type contour estimate, which shows that long contours are improbable. More precisely, we need to show that the probability for the occurrence of a contour $`\gamma `$ of size $`|\gamma |`$ is exponentially small in $`|\gamma |`$. By using again the chessboard estimate, it is possible to show that it is enough to obtain the desired estimate for just one single contour, $`\mathrm{\Gamma },`$ called the universal contour. This universal contour contains all bonds of $`\mathrm{\Lambda };`$ half of them are ordered, and the remaining half - disordered. In our case the event $`\mathrm{\Gamma }`$ happens iff all the bonds adjacent to sites $`x`$ with $`x_1+x_2=0\mathrm{mod}4`$ are ordered, while those adjacent to sites with $`x_1+x_2=2\mathrm{mod}4`$ are disordered. Since the size $`\left|\mathrm{\Gamma }\right|=\left|\mathrm{\Lambda }\right|,`$ we have to show that $$\mathrm{Pr}\left(\mathrm{\Gamma }\right)\mathrm{exp}\left\{c\left|\mathrm{\Lambda }\right|\right\}$$ (9) with the constant $`c`$ sufficiently large. (The concept of a ”universal contour” goes back to the pioneering paper by Fröhlich and Lieb . The remaining computations go just as in the proof for the large-$`q`$ Potts model, given in detail in or . This similarity with the Potts problem was already remarked upon in . The correspondence is that $`\epsilon `$, (and the same holds for $`\frac{1}{\sqrt{p}}`$), plays the role of the small parameter $`\frac{1}{q}`$. To estimate the probability of the universal contour, one has to integrate over all configurations such that the prescribed arrangement of ordered and disordered bonds occurs. To do it we observe that the total partition function of an $`N`$-by-$`N`$ square satisfies $$Z_N\mathrm{max}[1,(\frac{1}{2\pi }\epsilon exp2\beta )^{N^2}].$$ (10) Indeed, on the one hand we use that the potential is positive, and on the other hand we get a lower bound by taking the integral at each site over the interval $`[\frac{1}{2}\epsilon ,\frac{1}{2}\epsilon ]`$, so that each bond is ordered. The restricted partition function $`Z_\mathrm{\Gamma },`$ which is the integral over all configurations compatible with the universal contour satisfies $$Z_\mathrm{\Gamma }(2\epsilon )^{\frac{1}{4}N^2}\mathrm{exp}\left\{\frac{\beta N^2}{2}+O(N)\right\}.$$ (11) From these two estimates the bound $`\left(\text{9}\right)`$ follows. The final conclusion is now, that somewhere inside our temperature interval there is a value $`\beta _t,`$ at which a first-order transition happens between an ordered and a disordered Gibbs measure, as was first numerically found in . The value of $`\beta _t`$ is approximately given by $`2\beta _t=\mathrm{ln}\epsilon `$. The non-square-well model can be treated in a very similar way (see and also ). The ordered Gibbs measure has a polynomial spin-spin correlation decay of Kosterlitz-Thouless type. Generalisations: 1) The same method of proof works if either the spin dimensionality $`n`$ or the dimension $`d`$ of the lattice is larger than $`2`$ (or both). For the case of Heisenberg spins ($`n=3`$) in $`d=2`$, the first-order transition was first found numerically . In this case presumably both the low-temperature and the high-temperature phase have exponential decay of the spin-spin correlations. For the $`n\mathrm{}`$ spherical limit see also . For $`d3`$, the Mermin-Wagner theorem does not apply anymore, and the low-temperature phase now displays a spontaneous magnetisation. 2) In a small external field there still is a first-order transition between an ordered (strongly magnetised in the direction of the field) and a disordered (weakly magnetised in the direction of the field) phase, which now we expect to be both pure phases (extremal Gibbs measures) also in higher dimensions. 3) Instead of a single well, one can also consider potentials having the shape of repeated wells in wells (or a hat-in-a-hat-in-a-hat…), which give rise to possibly infinitely many transitions. A choice of such a Seuss () potential is $`U(x)=_n2^n1_{\epsilon _n}(x)`$ with $`\epsilon _n(=\epsilon _{n1}^3)=\epsilon ^{3^{n1}}`$, with the first $`\epsilon `$ small enough. Such transitions in which one keeps jumping in deeper and deeper wells, can occur either between nonmagnetised-nonmagnetised, magnetised-nonmagnetised or magnetised-magnetised Gibbs measures, and the nonmagnetised measures may display either exponential or Kosterlitz-Thouless decay. 4) Instead of ferromagnetic models the argument also works for nematic liquid crystal $`RP^n`$ models , in which one considers interactions of the form $`U(x)=x^{2p}`$ for which there are two minima on the interval $`[1,+1]`$. Here even for $`p=1`$, and $`n=3`$, first-order transitions were found numerically (see e.g. ) in $`d=3`$, whereas the occurrence of a transition in the limit $`n\mathrm{}`$ in $`d=2`$ has been a matter of controversy . Additional numerical references are mentioned in . Further models of this type, with a larger number of sharp minima, combined with a term causing a chiral symmetry breaking as in , are considered in . 5) Similarly as for the Potts gauge model of , we can prove the existence of a first-order transition in various nonlinear lattice gauge models with continuous symmetries. In some of these models first-order transitions were initially concluded on the basis of numerical data (see e.g. ). In we provide the first occasion where a first-order transition for a lattice gauge model in the presence of a continuous symmetry can be proven. 6) Instead of the above spin systems with a compact rotation symmetry, one can also consider continuous unbounded-spin systems which possess a non-compact symmetry (that is, they are “massless”). In this case the symmetry describes a shift in the height (average) of the spin. Again a similar construction works , now showing coexistence of “gradient Gibbs measures”. 7) For quantum spin systems the ingenious analysis of shows that again a first-order transition occurs, once the potential is sufficiently nonlinear and the spin number is large enough. ## 4 Conclusions and Comments In many cases, some of them also of direct physical interest, first-order transitions occur instead of the second-order transition, which a naive universality argument would predict. Such first-order transitions tend to occur more easily when either the nonlinearity parameter $`p`$, the spin-dimensionality $`n`$ (the spherical limit), or the dimension $`d`$ of the lattice (the mean-field limit, ) is large. In this paper we have presented a number of models where we can prove these transitions for sufficiently strong nonlinearities. Furthermore, if one adds an external field, the transition persists. Although the two phases on both sides of the transition can have different characters (breaking other symmetries, having polynomial or exponential decay of the spin-spin correlations, etc), it can be shown that no “intermediate” phases exist (there is a “forbidden gap” for the energy variable ). The method of Reflection Positivity we use here has the disadvantage that one is limited in the interactions one can take, in the sense that they need to be defined on the unit cube. As compared to the more robust Pirogov-Sinai contour methods, RP methods have the advantage, however, that they are more generally applicable in that we need no information about the phases on both sides of the transition. For Pirogov-Sinai methods to work, however, the phases need to be pure and will typically have some kind of exponential decay of correlations. On the other hand, with RP methods it has not been possible up to now to obtain results for surface or interface properties or about the completeness of the phase diagram. If one varies the nonlinearity parameter in $`d=2`$, one moves towards a critical point where the second-order transition is expected to have Ising characteristics, in higher dimensions varying the nonlinearity parameter will lead one to a tricitical point. Although our proofs require a fairly large nonlinearity (that is, a large value of $`p`$, or a small value of $`\epsilon `$), either by numerical methods or in the large-$`n`$ or large-$`d`$ limit we expect, and sometimes know, that the type of first-order transitions we studied will occur for much smaller values of $`p`$, $`n`$, and/or $`d`$, especially for the liquid-crystal models. *Acknowledgements* During this work we had very stimulating input from various colleagues, in particular L. Chayes, who developed independently a number of these ideas, and E. Domany and A. Schwimmer who urged us to also consider lattice gauge models. Some useful remarks on the manuscript by C. Külske are gratefully acknowledged.
warning/0506/math0506599.html
ar5iv
text
# Simplicity of QC(𝕊^𝑛) and LIP(𝕊^𝑛) ## 1 Introduction A group $`G`$ is simple if its only normal subgroups are itself and the trivial group. Let QC($`𝕊^n`$) and LIP($`𝕊^n`$) denote the orientation preserving quasiconformal and bilipschitz homeomorphisms of $`𝕊^n`$ respectively (We will implicitly assume throughout that the dimension $`n`$ is greater than 1). In this paper we prove the following statement, ###### Theorem 1.1 The groups QC($`𝕊^n`$) and LIP($`𝕊^n`$) are simple. Our proof falls into two stages. First, we show that the subgroups of QC($`𝕊^n`$) and LIP($`𝕊^n`$) which are generated by the elements supported on open balls are simple. In the second stage, we show that *every* element of QC($`𝕊^n`$) or LIP($`𝕊^n`$) can be written as a composition of elements which are all supported on open balls. The main part of the argument is Lemma 3.6 which proves that there are neighbourhoods of the identity in QC($`𝕊^n`$) and LIP($`𝕊^n`$) which satisfy this condition. Anderson proved that if $`G`$ is a group of homeomorphisms of a Hausdorff space $`X`$ and $`K`$ is a basis for $`X`$ satisfying certain conditions, then the group $`G_0`$ generated by the homeomorphisms which are supported on elements of $`K`$ is simple. In fact, he showed that if $`h`$ is any non trivial element of $`G`$, then every element of $`G_0`$ can be written as a product of four conjugates of $`h`$ and $`h^1`$. We demonstrate in detail that the group of quasiconformal homeomorphisms of $`𝕊^n`$ or the group of Lipschitz homeomorphisms of $`𝕊^n`$ and the basis consisting of open balls satisfy the necessary conditions to apply his argument and give the proof that these groups will therefore be simple. In his paper, Anderson also gave some conditions that the whole group $`G`$ will be generated by $`G_0`$. Namely that $`X`$ can be expressed as the union of two basis elements, and that $`G`$ has the ”Annulus property”. Here, $`G`$ satisfies the annulus property if for every $`gG`$ and $`kK`$ such that there exists $`k^{}K`$ with $`kg(k)k^{}`$, there exists $`g_0G_0`$ such that $`g`$ and $`g_0`$ agree on $`k`$. Now, $`𝕊^n`$ can obviously be written as the union of two open balls and in the process of proving Lemma 3.6 we show that there are neighbourhoods of $`id`$ in the groups QC($`𝕊^n`$) and LIP($`𝕊^n`$) which satisfy the annulus property. In these cases, this turns out to be enough. A homeomorphism $`h:XX`$ is called *stable* if it can be written as a finite composition of homeomorphisms, each of which is the identity on some open set. We say a group of homeomorphisms $`G`$ is *stable* if every element of $`G`$ is (See and for more information on stable homeomorphisms). Kirby showed that the group of homeomorphisms of the torus $`𝕋^n`$ is stable. He used this, along with an immersion of $`𝕋^n`$ minus a point into $`^n`$ to show that the group of homeomorphisms of $`^n`$ is locally contractible. Edwards and Kirby then went on to extend these results to arbitrary manifolds, in particular, showing that the homeomorphism group of a compact manifold $`M`$ is locally contractible. The stable homeomorphism conjecture says that the orientation preserving homeomorphisms of $`^n`$ are stable. This is known to be true for $`n4`$ whereas for $`n=4`$ this question is unsolved. A simple application of Anderson’s results concludes from the stable homeomorphism conjecture that the group of orientation preserving homeomorphisms of $`𝕊^n`$ is simple for $`n4`$. In general, if $`M`$ is a non compact manifold then its group of homeomorphisms will not be simple, as the subgroup consisting of elements with compact support forms a non-trivial normal subgroup. Likewise for manifolds with boundary, the subgroup consisting of elements which act as the identity in some neighbourhood of the boundary will also be a non-trivial normal subgroup. In the case of an arbitrary closed manifold the homeomorphism group has generally got more than one path component, with the component of the identity (which consists of all the stable homeomorphisms) forming a non-trivial normal subgroup. It is still an interesting question to ask whether these subgroups are themselves simple. For the case of diffeomorphisms, let $`M`$ be an $`n`$-dimensional smooth manifold. Define Diff$`(M,r)`$ to be the group of $`C^r`$ diffeomorphisms of $`M`$ which are isotopic to the identity through compactly supported $`C^r`$ isotopies (An isotopy has compact support if there is compact set outside of which the isotopy acts trivially). In and Mather proved that if $`\mathrm{}rn+1`$ then Diff$`(M,r)`$ is perfect (equal to its own commutator subgroup) with the case $`r=\mathrm{}`$ being proved by Thurston (See ). Now, in , Epstein showed that for certain groups of homeomorphisms, the commutator subgroup is simple. In particular, Diff$`(M,r)`$ is simple when $`\mathrm{}rn+1`$. Simplicity questions have also been addressed in the symplectic and volume preserving categories. Perhaps most notably Banyaga proved that for a closed symplectic manifold $`M`$ the group of Hamiltonian symplectomorphisms Ham$`(M,\omega )`$ is perfect and then used Epstein’s argument, which also applies in the symplectic case, to deduce that it is simple. On a non compact manifold, if one considers the component of the identity in the group of compactly supported symplectomorphisms $`\mathrm{Symp}_0^c(M,\omega )`$, then there exists a surjective homomorphism from $`\mathrm{Symp}_0^c(M,\omega )`$ to $``$. Banyaga proved that the kernel of this homomorphism is simple and also showed an analogous result in the volume preserving case. In general however, little is known about the normal subgroups of $`\mathrm{Symp}(M,\omega )`$. For example we currently do not know whether the group of symplectomorphisms of $`^{2n}`$ with its standard symplectic form is perfect. A stark contrast to the volume preserving case, where Mcduff has shown that $`\mathrm{Diff}_{\mathrm{vol}}(^n)`$ is perfect for every dimension $`n3`$. In Sullivan proved the stable homeomorphism conjecture and the annulus conjecture in the locally quasiconformal and Lipschitz categories in all dimensions. We go through his proof of the stable homeomorphism conjecture in detail, and use the result to show simplicity of QC($`𝕊^n`$) and LIP($`𝕊^n`$). I would like to thank Vladimir Markovic for introducing me to this problem and for many helpful discussions regarding this work. ## 2 QC and LIP homeomorphisms supported on open balls generate a simple group. In this section we start by introducing some notation and terminology and go on to prove that every element of QC($`𝕊^n`$) or LIP($`𝕊^n`$) which is supported on an open ball can be written as a composition of four conjugates of an arbitrary non-trivial element and its inverse. More importantly for us, this gives us the easy corollary that the subgroups of QC($`𝕊^n`$) and LIP($`𝕊^n`$) generated by those elements supported on open balls are simple. Let $`d_^n`$ denote the usual Euclidean distance on $`^n`$, we set $$𝕊^{n1}=\{x^n:d_^n(x,0)=1\}$$ to be the unit sphere in $`^n`$ and denote the standard spherical metric on $`𝕊^n`$ by $`d_{𝕊^n}`$. Let $$𝔹^n=\{x^n:d_^n(x,0)<1\}$$ denote the unit ball in $`^n`$, and let $`d_{𝔹^n}`$ denote its hyperbolic metric. We now define, $$B_X(c,ϵ)=\{xX:d_X(x,c)<ϵ\}$$ to be the open ball of radius $`ϵ>0`$ around $`cX`$, where $`X`$ is one of $`^n`$, $`𝕊^n`$ or $`𝔹^n`$ and $`d_X`$ is the appropriate metric. We also define a metric on the space of maps from $`𝕊^n`$ to itself by, $$\stackrel{~}{d}(f,g)=sup\{d(f(x),g(x)):x𝕊^n\}.$$ The group of Möbius transformations Möb($`𝕊^n`$) is the transformation group of $`𝕊^n`$ generated by inversions in ($`n1`$)-spheres and Möb$`{}_{}{}^{+}(𝕊^n)`$ is the index two subgroup consisting of those which preserve orientation. By identifying $`𝕊^n`$ with $`^n\{\mathrm{}\}`$ we can think of $`𝔹^n`$ as a subset of $`𝕊^n`$. If we define Möb($`𝔹^n`$)=$`\{f`$Möb$`(𝕊^n):f(𝔹^n)=𝔹^n\}`$ then Möb($`𝔹^n`$) is generated by inversions in ($`n1`$)-spheres orthogonal to $`𝕊^{n1}=𝔹^n`$ and is the isometry group of $`𝔹^n`$ with its hyperbolic metric. As before Möb$`{}_{}{}^{+}(𝔹^n)`$Möb$`(𝔹^n)`$ denotes those Möbius transformations which preserve orientation. The groups Möb$`(𝔹^n)`$ and Möb$`{}_{}{}^{+}(𝔹^n)`$ preserve $`𝕊^{n1}`$ and in the obvious way are isomorphic to Möb$`(𝕊^{n1})`$ and Möb$`{}_{}{}^{+}(𝕊^{n1})`$ respectively. ###### Definition 2.1 Let $`(X,d_X)`$ and $`(Y,d_Y)`$ be metric spaces and let $`f:XY`$ be a map between them. Then if there is a constant $`L>0`$ such that $$d_Y(f(x),f(y))Ld_X(x,y)$$ for every $`x,yX`$, then $`f`$ is $`L`$-*Lipschitz* . Furthermore, if we also have that $$d_Y(f(x),f(y))d_X(x,y)/L$$ for every $`x,yX`$ then $`f`$ is $`L`$-*bilipschitz*. ###### Definition 2.2 If $`f:(X,d_X)(Y,d_Y)`$ is such that every $`xX`$ has a neighbourhood $`U`$ such that the restriction of $`f`$ to $`U`$ is Lipschitz, then $`f`$ is *locally Lipschitz* (similarly for the definition of locally bilipschitz) ###### Remark 2.3 Metric spaces and locally Lipschitz maps form a category usually denoted by LIP. For this reason we will abbreviate locally Lipschitz to LIP. Moreover, if $`f:(X,d_X)(Y,d_Y)`$ is bijective and both $`f`$ and $`f^1`$ are LIP maps, then we say $`f`$ is a *lipeomorphism*. We write LIP($`X`$) to denote the space of lipeomorphisms from $`X`$ to itself. Since every LIP map with compact domain is Lipschitz, the space LIP($`𝕊^n`$) consists of all the bilipschitz homeomorphisms from $`𝕊^n`$ to itself (the point here is that we don’t have to say *locally* bilipschitz). It is a standard result that every diffeomorphism from $`𝕊^n`$ to itself lies in LIP($`𝕊^n`$). ###### Remark 2.4 We will only consider those homeomorphisms which preserve the orientation of $`𝕊^n`$. Let $`\mathrm{\Omega },\mathrm{\Omega }^{}`$ be subdomains of $`^n`$ and $`f:\mathrm{\Omega }\mathrm{\Omega }^{}`$ be a homeomorphism. Fix $`x_0\mathrm{\Omega }`$ and for small enough $`ϵ>0`$ consider the quantities $$D_l(ϵ)=inf\{d_^n(f(x_0),y):yf(B_^n(x_0,ϵ))\}$$ and $$D_u(ϵ)=sup\{d_^n(f(x_0),y):yf(B_^n(x_0,ϵ))\}$$ Let $`K_{x_0}(ϵ)=D_u(ϵ)/D_l(ϵ)1`$. If there exists $`K<\mathrm{}`$ such that $`K_{x_0}=limsup_{ϵ0}K_{x_0}(ϵ)<K`$ for every $`x_0\mathrm{\Omega }`$ then we say $`f`$ is $`K`$-quasiconformal. We say $`f`$ is quasiconformal if $`f`$ is $`K`$-quasiconformal for some $`K`$. If every point $`x_0\mathrm{\Omega }`$ has a neighbourhood $`U`$ such that $`f|_U`$ is $`K`$-quasiconformal, then we say $`f`$ is locally $`K`$-quasiconformal. Locally quasiconformal homeomorphisms between open subsets of $`^n`$ form a pseudogroup on $`^n`$ (See for the definition of a pseudogroup and more detail). In the usual way, this allows us to construct *quasiconformal manifolds*. These are manifolds whose transition maps are locally quasiconformal. Furthermore, if we have a homeomorphism $`f:MN`$ between two such manifolds, then we say it is locally quasiconformal if the following holds. For any pair of charts $`\varphi :M^n`$ and $`\psi :N^n`$ the composition $`\psi f\varphi ^1`$ is locally quasiconformal where defined. In particular, we let QC($`𝕊^n`$) denote the group of orientation preserving quasiconformal mappings from $`𝕊^n`$ to itself. Notice here that the standard smooth structure on $`𝕊^n`$ automatically induces a locally quasiconformal structure and as in the case of LIP($`𝕊^n`$), QC($`𝕊^n`$) contains all the orientation preserving diffeomorphisms of $`𝕊^n`$. ###### Remark 2.5 The space of lipeomorphisms between open subsets of $`^n`$ also forms a pseudogroup. So we could equally well have defined LIP($`𝕊^n`$) to be the group of orientation preserving bilipschitz homeomorphisms of $`𝕊^n`$ where its Lipschitz structure is induced from the standard smooth structure. Throughout the rest of the paper we take G($`𝕊^n`$) to denote either LIP($`𝕊^n`$) or QC($`𝕊^n`$). Take $``$ to be the basis for the induced topology on $`𝕊^n`$ consisting of the open balls $`\{B_{𝕊^n}(c,ϵ):c𝕊^n,ϵ(0,\pi )\}`$. This ensures that for all $`B`$, $`𝕊^nB`$ is homeomorphic to $`\overline{𝔹}^n`$ by excluding the possibility that $`B=𝕊^n`$ or $`𝕊^n`$ minus a point. We do this to ensure that given two elements $`B_1,B_2`$ we can find $`gG(𝕊^n)`$ such that $`g(B_1)=B_2`$. If $`g\mathrm{G}(𝕊^n)`$ is not the identity, then we say $`g`$ is *supported* on $`B`$ if $`g`$ is the identity outside $`B`$. Consequently, if $`B_1B_2`$ and $`g\mathrm{G}(𝕊^n)`$ is supported on $`B_1`$ then $`g`$ is also supported on $`B_2`$. We set $`\mathrm{G}_0(𝕊^n)`$ to be those elements of $`\mathrm{G}(𝕊^n)`$ supported on elements of $``$. We have already mentioned that for all $`B_1,B_2`$ there exists $`g\mathrm{G}(𝕊^n)`$ with $`g(B_1)=B_2`$. In particular, if $`B_1,B_2`$ and $`g\mathrm{G}_0(𝕊^n)`$ is an element supported on $`B_1`$, then $`g`$ is conjugate to an element of $`\mathrm{G}_0(𝕊^n)`$ supported on $`B_2`$. Furthermore, we can choose $`g`$ to lie in Möb$`{}_{}{}^{+}(𝕊^n)\mathrm{G}(𝕊^n)`$ so that if $`B`$ then $`g(B)`$ will also lie in $``$. ###### Definition 2.6 A pair $`(\{B_i\}_{i0},\rho )`$ consisting of a sequence $`\{B_i\}_{i0}`$ of elements of $``$, and $`\rho \mathrm{G}_0(𝕊^n)`$ will be called a $`\sigma `$-sequence if, 1. $`B_iB_j=\varphi `$ if $`ij`$ 2. There exists $`x_0𝕊^n`$ such that every neighbourhood of $`x_0`$ contains all but finitely many of the $`B_i`$ 3. $`\rho (B_i)=B_{i+1}`$ This means that if $`\rho `$ is supported on $`B`$ then $`_{i0}B_iB`$ (since condition 1 means that $`\rho `$ cannot be the identity anywhere in $`_{i0}B_i`$). If $`(\{B_i\}_{i0},\rho )`$ is a sigma sequence, we will say it is supported on $`B`$ if $`\rho `$ is. ###### Example 2.7 We now use the fact that Möb$`{}_{}{}^{+}(𝕊^n)\mathrm{G}(𝕊^n)`$ to construct a $`\sigma `$-sequence explicitly. Consider $`𝕊^n`$ as $`^n\{\mathrm{}\}`$ then the Möbius transformation, $$f(\text{x})=\frac{\text{x}}{r^2},f(\mathrm{})=\mathrm{}$$ with $`r\{0\}`$ is the time $`r^2`$ flow of the complete vector field, $$𝒳(x_1,\mathrm{},x_n)=\mathrm{\Sigma }_ix_i\frac{}{x_i}$$ Let $`\phi :𝕊^n`$ be a smooth function which is identically one on $`B_{𝕊^n}(0,1)`$ and zero outside $`B_{𝕊^n}(0,1+ϵ)`$ for some $`0<ϵ<1`$ and take $`\rho `$ to be the time $`r^2`$ flow of $`\phi 𝒳`$. Then $`\rho `$ will be a smooth map which is the same as $`f`$ on $`B_{𝕊^n}(0,1)`$ and the identity outside $`B_{𝕊^n}(0,1+ϵ)`$ and so is in $`\mathrm{G}_0(𝕊^n)`$. Now take $`B`$ with $`BB_{𝕊^n}(0,1)`$ and $`\rho (B)B=\varphi `$ and define $`\{B_i\}=\rho ^i(B)`$ $`i0`$, these will all be in $``$ since $`\rho `$ is conformal on $`B_{𝕊^n}(0,1)`$ and $`(\{B_i\},\rho )`$ will be a $`\sigma `$-sequence supported on $`B_{𝕊^n}(0,1+ϵ)`$ with $`x_0=0`$. Take any $`B`$, then since there exists $`g\mathrm{G}(𝕊^n)`$ with $`g(B_{𝕊^n}(0,1))=B`$, we have that $`(\{g(B_i)\},g\rho g^1)`$ is a $`\sigma `$-sequence supported on $`g(B_{𝕊^n}(0,1+ϵ))`$. Let $`(\{B_i\},\rho )`$ be a $`\sigma `$-sequence supported on $`B`$, for each $`i`$ take $`g_i\mathrm{G}_0(𝕊^n)`$ which is supported on $`B_i`$. We would like to be able to say that $`_{i0}g_i=\mathrm{}g_1g_0`$ is also in $`\mathrm{G}_0(𝕊^n)`$. If $`g_i\mathrm{QC}_0(𝕊^n)`$ for each $`i`$ and there exists K $`1`$ such that the $`g_i`$ are all K-quasiconformal, then the composition $`_{i0}g_i`$ is also K-quasiconformal and hence in $`\mathrm{QC}_0(𝕊^n)`$. In particular, if $`\rho |_{B_i}`$ is conformal (this was the case in the example we constructed above) and we take $`g_0`$ supported on $`B_0`$ which is K-quasiconformal. Then the conjugates $`\rho ^ig_0\rho ^i`$ will also be K-quasiconformal and supported on $`B_i`$. Hence, the composition $`_{i0}\rho ^ig_0\rho ^i`$ will also be K-quasiconformal, and thus lie in $`\mathrm{QC}_0(𝕊^n)`$. Similarly, if for each $`i`$, $`g_i\mathrm{LIP}_0(𝕊^n)`$ is supported on $`B_i`$ and there exists $`L1`$ such that the Lipschitz constants of the $`g_i`$ are bounded by $`L`$, then the composition $`_{i0}g_i`$ is $`L`$-Lipschitz and hence in $`\mathrm{LIP}_0(𝕊^n)`$. In particular, if we take $`g_0`$ supported on $`B_0`$ which is $`L`$-Lipschitz. Then again using $`\rho `$ from the example above, the conjugates $`\rho ^ig_0\rho ^i`$ will also be $`L`$-Lipschitz and supported on $`B_i`$. Hence, the composition $`_{i0}\rho ^ig_0\rho ^i`$ will also be $`L`$-Lipschitz, and thus lie in $`\mathrm{LIP}_0(𝕊^n)`$. We will use these observations to prove the following theorem. ###### Theorem 2.8 Given a non-trivial $`h\mathrm{G}(𝕊^n)`$ , any element $`g\mathrm{G}_0(𝕊^n)`$ is the product of four conjugates of $`h`$ and $`h^1`$. Proof. As $`h`$ is not the identity, we can find $`B`$ with $`h^1(B)B=\varphi `$. Let $`(\{B_i\},\rho )`$ denote a $`\sigma `$-sequence with $`_iB_iB`$ obtained as in Example 2.7, so that $`\rho |_B`$ is conformal (Note that $`\rho `$ will be supported on some larger ball containing $`B`$). Since every element of $`\mathrm{G}_0(𝕊^n)`$ can be conjugated in $`\mathrm{G}(𝕊^n)`$ to one supported on $`B_0`$ it suffices to show that any $`g`$ supported on $`B_0`$ is the product of four conjugates of $`h`$ and $`h^1`$. So let $`g\mathrm{G}_0(𝕊^n)`$ be supported on $`B_0`$, then the map, $$f=\underset{n0}{}\rho ^ng\rho ^n$$ is an element of $`\mathrm{G}_0(𝕊^n)`$ supported on $`_{i0}B_iB`$. Since the conjugate $`h^1f^1h`$ is then supported on $`h^1(B)`$ and $`h^1(B)B=\varphi `$, we can think of the commutator, $$f_1:=h^1f^1hf=(h^1f^1h)f$$ as two actions on disjoint sets. The same can be said about, $$f_2:=\rho h^1fhf^1\rho ^1=(\rho h^1fh\rho ^1)(\rho f^1\rho ^1)$$ here $`\rho f^1\rho ^1`$ is supported on $`B`$ whereas $`\rho h^1fh\rho ^1`$ is supported on $`h^1(B)`$. So $`f_2f_1`$ can be examined in terms of its action on $`B`$ and $`h^1(B)`$ respectively. Doing this, one can easily see that $`f_2f_1`$ is the identity on $`h^1(B)`$ and acts as $`g`$ on $`B`$. Since $`g`$ is supported on $`B_0B`$, $`g=f_2f_1`$, but we can write, $$g=f_2f_1=(\rho h^1\rho ^1)(\rho fhf^1\rho ^1)(h^1)(f^1hf)$$ as a product of four conjugates of $`h`$ and $`h^1`$. margin: $`\mathrm{}`$ ###### Corollary 2.9 The subgroup of $`\mathrm{G}(𝕊^n)`$ generated by $`\mathrm{G}_0(𝕊^n)`$ is simple. Proof. Denote by $`\mathrm{G}(\mathrm{G}_0(𝕊^n))`$ the subgroup of $`\mathrm{G}(𝕊^n)`$ generated by $`\mathrm{G}_0(𝕊^n)`$ and let $`H`$ be a non-trivial normal subgroup. We need to show that $`H=\mathrm{G}(\mathrm{G}_0(𝕊^n))`$. So let $`g\mathrm{G}(\mathrm{G}_0(𝕊^n))`$ and take $`hH`$ which is not the identity. We can write $`g`$ as a finite composition $`g_n\mathrm{}g_1`$ where each $`g_i\mathrm{G}_0`$ for $`i=1\mathrm{}n`$. Then using Theorem 2.8 we can write each $`g_i`$ as a product of four conjugates of $`h`$ and $`h^1`$, but since $`H`$ is normal this implies that each $`g_iH`$ and thus $`gH`$ which gives us the result. margin: $`\mathrm{}`$ It now remains to show that $`\mathrm{G}(𝕊^n)`$ is equal to $`G(\mathrm{G}_0(𝕊^n))`$. We show this in the following section. ## 3 Sullivan groups In this section we introduce the concept of a Sullivan group, this is a discrete group of hyperbolic isometries which gives rise to a hyperbolic n-manifold with the property that the complement of any point can be LIP-immersed into $`^n`$. We use this along with two versions of the famous Schoenflies Theorem in the LIP category to prove the main result, Lemma 3.6, which states that there is a neighbourhood of the identity in G($`𝕊^n`$) which is a subset of $`G(\mathrm{G}_0(𝕊^n))`$. Using this, we prove that $`G(\mathrm{G}_0(𝕊^n))`$ is equal to G($`𝕊^n`$) which then must be simple. If $`(X,d_X)`$ and $`(Y,d_Y)`$ are metric spaces and $`f:(X,d_X)(Y,d_Y)`$ is a lipeomorphism onto its image (and hence injective) then we say $`f`$ is a *LIP embedding* of $`X`$ in $`Y`$. If every $`xX`$ has a neighbourhood $`U`$ such that $`f|_U:UY`$ is a LIP embedding then we say $`f`$ is a *LIP immersion*. Let $`\mathrm{\Gamma }`$Möb$`{}_{}{}^{+}(𝔹^n)`$ be a discrete group of Möbius transformations which acts freely, then the quotient space $`Q=𝔹^n/\mathrm{\Gamma }`$ is a smooth Hausdorff manifold and the natural map $`\pi :𝔹^nQ`$ is a covering map. Note that since $`\mathrm{\Gamma }`$ acts properly and $`𝔹^n`$ is locally compact, the property of $`\mathrm{\Gamma }`$ being discrete is equivalent to it acting properly discontinuously. $`Q`$ inherits a hyperbolic metric from $`𝔹^n`$ defined in the usual way so that $`\pi `$ becomes a local isometry and we write $`B_Q(q,ϵ)`$ to denote the open ball of radius $`ϵ`$ around $`qQ`$ in this metric. ###### Definition 3.1 A group $`\mathrm{\Gamma }`$ of Möbius transformations acting on $`𝔹^n`$ is a Sullivan Group if it satisfies the following; 1. $`\mathrm{\Gamma }`$ is discrete 2. $`\mathrm{\Gamma }`$ acts freely 3. $`Q=𝔹^n/\mathrm{\Gamma }`$ is closed (compact and without boundary) 4. For every $`qQ`$ there exists a LIP immersion $`\iota :Qq^n`$ If $`\mathrm{\Gamma }`$ is a Sullivan group we call the quotient space $`Q=𝔹^n/\mathrm{\Gamma }`$ a *Sullivan manifold*. We have the following theorem from . ###### Theorem 3.2 There exists a Sullivan manifold of every dimension $`n2`$. Sullivan’s proof of this theorem is deep and we take it on faith here. As an easy example though, in dimension 2 any closed surface of genus $`𝐠2`$ can be constructed as a quotient of $`𝔹^n`$ by a group of Möbius transformations $`\mathrm{\Gamma }`$ which is discrete and acts freely. We claim that all of these groups are Sullivan groups. To see this it remains to show condition 4, that for every $`qQ`$ there exists a LIP immersion $`\iota :Qq^n`$. So take $`qQ`$ and remove the ball $`B_Q(q,ϵ)`$. The resulting surface is lipeomorphic (diffeomorphic in fact) to a disc with $`2𝐠`$ strips attached (see Figure 1), which can then be LIP-immersed into $`^n`$. Letting $`ϵ`$ tend to 0 then gives the required map (note however that the bilipschitz constants will become arbitrarily large near $`q`$). We will also need the following two theorems which we state without proof. ###### Theorem 3.3 (Lipschitz Schoenflies Theorem ) Suppose that $`0<a<1`$ and that $`f`$ is a LIP embedding of the annulus $`A=\overline{B}^nB_^n(0,a)`$ into $`𝕊^n`$. Then $`f|_{S^{n1}}`$ can be extended to a LIP embedding $`f^{}:\overline{B}^n𝕊^n`$. ###### Theorem 3.4 (Quantitative Schoenflies Theorem ) Let $`A=\overline{B}^nB_^n(0,1/2)`$, $`𝒩=\{f:A^n:sup\{d(x,f(x)):xA\}<1/30\}`$ and $`=\{f:\overline{B}^n^n\}`$ then there exists a continuous map $`\varphi :𝒩`$ a constant $`a_0`$ and constants $`b_n`$ which depend only on $`n`$ such that: 1. $`\varphi (f)|_{S^{n1}}=f|_{S^{n1}}`$ for all $`f𝒩`$ 2. $`\varphi (id)=id`$ 3. If $`f`$ is $`K`$-quasiconformal, then $`\varphi (f)`$ is $`a_nK^3`$-quasiconformal 4. If f is locally L-bilipschitz, $`\varphi (f)`$ is locally $`a_0L^3`$-bilipschitz 5. If f is L-bilipschitz, $`\varphi (f)`$ is $`a_0L^3`$-bilipschitz ###### Remark 3.5 In the definition of $``$ and $`𝒩`$ above we only consider $`f:A^n`$ which are embeddings. ###### Lemma 3.6 There exists $`ϵ>0`$ so that $$\{fG(𝕊^n):\stackrel{~}{d}(id,f)<ϵ\}G(G_0(𝕊^n))$$ (Recall that $`\stackrel{~}{d}(f,g)=sup\{d_{𝕊^n}(f(x),g(x)):x𝕊^n\}`$). Proof. Take $`f\mathrm{G}(𝕊^n)`$. Under the assumption that $`f`$ is close to $`id`$ we shall construct $`g\mathrm{G}_0(𝕊^n)`$ which agrees with $`f`$ on some open ball $`B`$. This will mean that $`g^1f`$ will be supported in the interior of $`B^c`$, and since $`int(B^c)`$, $`g^1f`$ will be in $`\mathrm{G}_0(𝕊^n)`$. As a result $`f=g(g^1f)`$ will be an element of $`G(\mathrm{G}_0(𝕊^n))`$ as required and the proof will be complete. Begin by using Theorem 3.2 to find a Sullivan group $`\mathrm{\Gamma }`$, giving rise to a Sullivan manifold $`Q=𝔹^n/\mathrm{\Gamma }`$ of dimension $`n`$. Let $`\pi :𝔹^nQ`$ denote the standard projection, then we have the following: For every $`qQ`$, there exists $`r(q)>0`$ such that for any $`p\pi ^1(q)`$ there exists a neighborhood $`U`$ of $`\overline{B}_{𝔹^n}(r(q),p)`$ such that the restriction $`\pi |_U`$ is an isometry. Since $`Q`$ is compact we can fix $`r>0`$ which works for every $`qQ`$ and define $`D=B_{𝔹^n}(0,r)𝔹^n`$. Consider the restriction of $`f`$ to the closed unit ball $`\overline{𝔹}^n𝕊^n`$, which we denote by $`f_0`$. We shall construct $`g:𝕊^n𝕊^n`$ which is supported on $`𝔹^n`$ and agrees with $`f`$ on $`D`$. It is thus necessary that we take $`f`$ sufficiently close to the identity so as to ensure that $`f_0(D)𝔹^n`$. Choose $`qQ\pi (\overline{D})`$ and use condition 4 of Definition 3.1 to find a LIP immersion $`\iota :QqB_^n(0,0.9)`$. We can arrange that $`\iota \pi =id`$ in a neighbourhood of $`\overline{D}`$. To see this, let $`U`$ be a neighbourhood of $`\overline{D}`$ on which $`\iota \pi `$ is a LIP-embedding. Let $`B_1,B_2𝔹^n`$ denote open balls centred at $`0`$ which satisfy $`\overline{D}B_1\overline{B}_1B_2U`$. Apply Theorem 3.3 to the map $`(\iota \pi )^1`$ on the annulus $`B_2B_1`$ to find a LIP-embedding $`h^{}:𝕊^nB_1𝕊^n`$ which agrees with $`(\iota \pi )^1`$ on $`B_1`$. Now define $`h:𝕊^n𝕊^n`$ by, $$h(x)=\{\begin{array}{cc}(\iota \pi )^1(x)\hfill & \text{if }xB_1\hfill \\ h^{}(x)\hfill & \text{if }x𝕊^nB_1\hfill \end{array}$$ then $`h`$ is a lipeomorphism which agrees with $`(\iota \pi )^1`$ in a neighbourhood of $`\overline{D}`$. By post-composing $`\iota `$ with $`h`$ we will get a LIP immersion of $`Q`$ into $`𝕊^n`$ whose image is contained in some element of $``$ which contains $`\overline{D}`$. We may now use a diffeomorphism which is the identity on some neighbourhood of $`\overline{D}`$ to map $`h(B_^n(0,0.9))`$ back inside $`B_^n(0,0.9)`$. This will give us a new LIP immersion $`\iota ^{}:QqB_^n(0,0.9)`$ with the additional property that $`\iota ^{}\pi =id`$ in a neighbourhood of $`\overline{D}`$. We now assume that $`\iota \pi =id`$ in a neighbourhood of $`\overline{D}`$. Choose $`t>0`$ sufficiently small so that $`3t<r`$ and for $`i=1,2,3`$ the balls $`D_i=B_Q(q,it)Q`$, satisfy $`D_i\pi (\overline{D})=\varphi `$. We will construct $`g`$ by lifting maps as shown. $$\begin{array}{cccc}\hfill g:& 𝕊^n& & 𝕊^n\\ & & & \\ \hfill \stackrel{~}{f}_2:& 𝔹^n& & 𝔹^n\\ & \pi & & \pi \\ \hfill f_2:& Q& & Q\\ & & & \\ \hfill f_1:& QD_2& & QD_1\\ & \iota & & \iota \\ \hfill f_0:& \overline{𝔹}^n& & 𝕊^n\end{array}$$ Firstly we construct $`f_1:QD_2QD_1`$, this is a natural lift of $`f_0`$ constructed using the LIP-immersion $`\iota `$. In order to define $`f_1`$ we require that $`f`$ is sufficiently close to $`id`$ (i.e. that $`f_0`$ is close to the inclusion $`\eta :\overline{B}^n^n\{\mathrm{}\}`$). This requirement is to ensure that $`f_0\iota (QD_2)`$ is contained within the image of $`\iota (QD_1)`$, a necessary condition for the map $`f_1`$ to exist. To define $`f_1`$, we first construct a finite open cover $`U_i`$ for $`i=1,\mathrm{},k`$ of $`QD_2`$, such that the restriction $`\iota |_{U_i}`$ is a LIP embedding for every $`i`$. Now construct another finite cover $`W_i`$ for $`i=1,\mathrm{},k`$, this time consisting of closed and hence compact subsets of $`QD_1`$ such that $`W_iU_i`$. By taking $`f`$ closer to $`id`$ if necessary we will have $`f_0(\iota (W_i))\iota (U_i)`$ so we can define $`f_1`$ on each $`W_i`$ by $`f_1|_{W_i}=(\iota |_{U_i})^1f_0(\iota |_{W_i})`$. This definition makes sense since $`\iota `$ is a LIP-embedding on each $`U_i`$ and hence has a well defined inverse on its image. Now by choosing $`f`$ to lie in an even smaller neighbourhood of the identity if necessary, we can apply the quantitative Schoenflies theorem 3.4 to the restriction of $`f_1`$ to $`\overline{D}_3D_2`$ to construct a map $`f_2:QQ`$ which agrees with $`f_1`$ on $`QD_3`$. Now we can lift $`f_2`$ to $`\stackrel{~}{f}_2:𝔹^n𝔹^n`$ in the usual way so that $`\pi \stackrel{~}{f}=f\pi `$. This lifting is unique if we specify $`\stackrel{~}{f}_2(0)\pi ^1f_2\pi (0)`$, but we have only one choice since we want $`\stackrel{~}{f}_2|_D=f_0|_D`$, namely we set $`f_0(0)=\stackrel{~}{f}_2(0)`$. Now, if we choose $`f`$ to be still closer to the identity, then since $`\iota (Qq)`$ is compactly contained in $`𝔹^n`$, $`f_1`$ will move closer to the identity map of $`QD_2`$ (Here we are using the hyperbolic metric on $`Q`$ to determine this distance). Consequently the restriction of $`f_1`$ to $`D_3`$ will also become closer to the identity (we used this to apply the quantitative Schoenflies theorem 3.4). Since the map $`\varphi `$ in theorem 3.4 is continuous and it fixes $`id`$, $`\varphi (f_1|_{\overline{D}_3D_2})`$ will also move closer to the identity. As a result we are making $`f_2`$ close to the identity in the hyperbolic metric on $`Q`$. If the dimension of $`Q`$ is 3 or more then by taking $`f_2`$ close enough to the identity we can deduce that $`f_2`$ is homotopic to the identity by applying Mostow rigidity. In dimension 2 however, we will show that $`f_2`$ will be homotopic to the identity by showing that it acts as the identity on $`\pi _1`$. To see this, take a set of generators $`\stackrel{~}{\gamma }_1\mathrm{}\stackrel{~}{\gamma }_m`$ for $`\pi _1`$ (this will be a finite set as $`Q`$ is closed) and take a representative $`\gamma _i`$ of each. Then for each $`\gamma _i`$ there exists $`ϵ_i`$ such that the $`ϵ_i`$-neigbourhood of $`\gamma _i`$ is homotopy equivalent to $`\gamma _i`$. The same will then be true of $`ϵ^{}=`$min$`\{ϵ_i:i=1,\mathrm{},m\}`$. One can now see that if $`f_2`$ is uniformly $`ϵ^{}`$ close to the identity then $`f_2(\gamma _i)`$ will be homotopic to $`\gamma _i`$ for each $`i`$. Consequently, it will induce the identity on $`\pi _1`$. The homotopy from $`f_2`$ to the identity then lifts to show that $`\stackrel{~}{f}_2`$ also homotopic to the identity. Now, since $`Q`$ is compact there exists a compact region $`K𝔹^n`$ such that $`\pi (K)=Q`$. Furthermore, the lifted homotopy commutes with the action of $`\mathrm{\Gamma }`$ and so is completely determined by its behaviour on $`K`$. Now, each point in $`K`$ can only be moved a finite hyperbolic distance by the homotopy and hence the same applies to $`\stackrel{~}{f}_2`$. Since $`\mathrm{\Gamma }`$ acts by isometries this applies to every point in $`𝔹^n`$ allowing us to deduce that $`\stackrel{~}{f}_2`$ is a bounded hyperbolic distance from the identity. We now use the fact that a lipeomorphism or K-quasiconformal homeomorphism of hyperbolic space which is a bounded hyperbolic distance from the identity will determine a map of the same class from $`\overline{B}^n`$ to itself which is the identity on $`𝔹^n`$. This means that we can extend $`\stackrel{~}{f}_2`$ to a map $`g:𝕊^n𝕊^n`$ which agrees with $`\stackrel{~}{f}_2`$ on $`𝔹^n`$ and is the identity otherwise. From the construction of $`g`$ we have that $`f_0\iota \pi =\iota \pi g`$ where defined (in particular this applies in a neighbourhood of $`D`$) and $`\iota \pi =id`$ in a neighbourhood of $`D`$, consequently $`f=g`$ on $`D`$ as we required. It remains to check that $`g`$ is in fact an element of $`G_0(𝕊^n)`$. Suppose that $`f`$ was quasiconformal and let $`E_1=\pi ^1(\overline{D}_3)`$ and $`E_2=\pi ^1(QD_3)`$. Then $`E_1`$ consists of countably many components $`E_1^i`$. Consider one such component $`E_1^1`$, then the dilatations of $`\stackrel{~}{f}_2`$ on $`E_1^1`$ are determined by the behaviour of $`f_2`$ on $`\overline{D}_3`$. Since $`f_2|_{\overline{D}_3}`$ was defined by the quantitative Schoenflies theorem it is quasiconformal, and hence $`\stackrel{~}{f}_2|E_1^1`$ will be too. Furthermore if $`ij`$ there exists $`\gamma \mathrm{\Gamma }`$ such that $`\stackrel{~}{f}_2|_{E_1^i}=\gamma \stackrel{~}{f}_2\gamma ^1|_{E_1^j}`$ since $`\gamma `$ is conformal $`g|_{E_1}`$ is quasiconformal. Now since $`Q`$ is compact, there exists an open set $`U𝔹^n`$ whose closure in $`𝔹^n`$ is compact and such that $`\pi (U)=Q`$, then $`\iota \pi `$ is locally bilipschitz on $`U\pi ^1(\overline{D}_1)`$. By construction, $`\iota \pi \stackrel{~}{f}_2=f_0\pi \iota `$ in $`E_2`$ and hence $`\stackrel{~}{f}_2|_{E_2U}`$ is quasiconformal. Since $`\gamma \stackrel{~}{f}_2=\stackrel{~}{f}_2\gamma `$ for all $`\gamma \mathrm{\Gamma }`$, we get $`\stackrel{~}{f}_2|_{E_2}`$ is quasiconformal. It now follows (see Theorem 35.1) that $`g`$ is quasiconformal. If $`f`$ was a lipeomorphism of $`𝕊^n`$ then the same argument as above will allow us to deduce that $`\stackrel{~}{f}_2`$ is a lipeomorphism of $`𝔹^n`$ with its hyperbolic metric. This, along with the fact that $`\stackrel{~}{f}_2`$ is close to the identity, implies that $`\stackrel{~}{f}_2`$ is a lipeomorphism of $`𝔹^n`$ with the Euclidean metric and hence $`g`$ LIP($`𝕊^n`$). margin: $`\mathrm{}`$ ###### Theorem 3.7 $`\mathrm{G}(𝕊^n)=G(\mathrm{G}_0(𝕊^n))`$ Proof. $`G(\mathrm{G}_0(𝕊^n))`$ is a subgroup of the topological group $`\mathrm{G}(𝕊^n)`$ which by Lemma 3.6 contains a neighbourhood of the identity and consequently we can translate this neighbourhood to show that $`G(\mathrm{G}_0(𝕊^n))`$ is open in $`\mathrm{G}(𝕊^n)`$. As $`G(\mathrm{G}_0(𝕊^n))`$ is an open subgroup, it will be closed, as its cosets will also be open. Consequently, it suffices to show that $`\mathrm{G}(𝕊^n)`$ is connected. So take $`f\mathrm{G}(𝕊^n)`$, we can easily construct a path in $`\mathrm{G}(𝕊^n)`$ which joins $`f`$ to an element with a fixed point, so assume $`f`$ has a fixed point $`x𝕊^n`$. Stereographic projection from $`x`$ allows us to identify $`𝕊^n\{x\}`$ with $`^n`$. We can consequently think of $`f`$ as a G-homeomorphism of $`^n`$, in particular there will exist $`x_0^n`$ at which $`f`$ is differentiable with positive Jacobian. Furthermore, we may assume that $`x_0=0=f(x_0)`$. Set, $`g_t(x)=f(tx)/t`$ for $`t>0`$ and $`g_0=f^{}(0)`$, then the path $`g_t`$ joins $`f^{}(0)`$ to $`f`$ and since $`det(f^{}(0))>0`$ we can join $`f^{}(0)`$ to the identity. We can extend all of these maps to $`𝕊^n`$ by leaving $`x`$ invariant giving us a path in $`\mathrm{G}(𝕊^n)`$ from $`f`$ to the identity. This shows that $`\mathrm{G}(𝕊^n)`$ is path connected and hence connected. margin: $`\mathrm{}`$
warning/0506/gr-qc0506131.html
ar5iv
text
# On the nature of initial singularities for solutions of the Einstein-Vlasov-scalar field system with surface symmetry ## 1 Introduction In the mathematical study of inhomogeneous cosmological solutions of the Einstein equations coupled to the Vlasov equation and a linear scalar field was begun. A local existence theorem and continuation criteria were proved for solutions with surface symmetry (i.e., spherical, plane or hyperbolic symmetry). In this was used to prove a global existence theorem in the expanding direction. In the case of plane symmetry with only a scalar field it was shown that the solutions are future geodesically complete. In the present paper we turn to the study of the past time direction, i.e., the approach to the initial singularity. In the case where the matter is given by the Vlasov equation alone certain results on the initial singularity have been obtained by Rein . He showed that if the maximum momentum of the particles remains bounded on any interval where the solution exists then, for spherical or plane symmetry, the solutions can be extended up to $`t=0`$. In the case of hyperbolic symmetry he showed that the corresponding result holds provided the metric function $`\mu `$ (defined below) is initially negative. Weaver showed that the solution always exists up to $`t=0`$ in the case of plane symmetry. This is a special case of a result she proved for solutions of the Einstein-Vlasov system with $`T^2`$-symmetry. Tchapnda extended the result from plane symmetry to spherical symmetry and to hyperbolic symmetry under the assumption that $`\mu `$ is initially negative or the assumption that $`\mu `$ is bounded on any interval where the solution exists. His results also allow for a negative cosmological constant. In the case where the matter is given by a scalar field alone existence up to $`t=0`$ for solutions with plane symmetry follows from the work of . In that case it is only necessary to solve a linear hyperbolic equation which is identical to the essential field equation in polarized Gowdy spacetimes. Using results of it was shown that the initial singularity is a curvature singularity (the Kretschmann scalar blows up uniformly there) and a crushing singularity (the mean curvature of the hypersurfaces of constant $`t`$ blows up uniformly as $`t0`$). For matter described by the Vlasov equation alone the results already mentioned can be combined with theorems in to show that $`t=0`$ is a crushing singularity where the Kretschmann scalar blows up uniformly. For an open set of initial data belonging to a restricted class the asymptotic behaviour can be described more precisely. For convenience these will be referred to as small data. The generalized Kasner exponents converge uniformly to the values $`(2/3,2/3,1/3)`$. This means that in a certain sense the solution is approximated near the singularity by a Kasner solution with those particular values of the Kasner exponents. In this paper some of the above results will be generalized to the case where both Vlasov matter and a scalar field are present. It is shown that the solutions exist up to $`t=0`$ for spherical and plane symmetry and in the case of hyperbolic symmetry when $`\mu `$ satisfies the restrictions mentioned above. The Kretschmann scalar and the mean curvature of the hypersurfaces of constant $`t`$ blow up uniformly as $`t0`$. Under the assumption that the maximum momentum of any particle in the radial direction decays like a positive power of $`t`$ as $`t0`$ it is possible to obtain further interesting estimates. Unfortunately we did not succeed in generalizing the small data results for the Vlasov equation alone to the present case. For a scalar field alone these estimates allow the detailed asymptotics to be determined for all solutions. The generalized Kasner exponents converge to the values $`((1a(r))/2,(1a(r))/2,a(r))`$ for a continuous function $`a(r)`$. It may be noted that these solutions have the same type of singularity as those constructed from data on the singularity in . Let us recall the formulation of the Einstein-Vlasov-scalar field system as shown in and . We consider a four-dimensional spacetime manifold $`M`$, with local coordinates $`(x^\alpha )=(t,x^i)`$ on which $`x^0=t`$ denotes the time and $`(x^i)`$ the space coordinates. Greek indices always run from $`0`$ to $`3`$, and Latin ones from $`1`$ to $`3`$. On $`M`$, a Lorentzian metric $`g`$ is given with signature $`(,+,+,+)`$. We consider a self-gravitating collisionless gas and restrict ourselves to the case where all particles have the same rest mass, normalized to $`1`$, and move forward in time. We denote by $`(p^\alpha )`$ the momenta of the particles. The conservation of the quantity $`g_{\alpha \beta }p^\alpha p^\beta `$ requires that the phase space of the particle is the seven-dimensional submanifold $$PM=\{g_{\alpha \beta }p^\alpha p^\beta =1;p^0>0\}$$ of $`TM`$ which is coordinatized by $`(t,x^i,p^i)`$. If the coordinates are such that the components $`g_{0i}`$ vanish then the component $`p^0`$ is expressed by the other coordinates via $$p^0=\sqrt{g^{00}}\sqrt{1+g_{ij}p^ip^j}$$ The distribution function of the particles is a non-negative real-valued function denoted by $`f`$, that is defined on $`PM`$. In addition we consider a massless scalar field $`\varphi `$ which is a real-valued function on $`M`$. The Einstein-Vlasov-scalar field system now reads: $$_tf+\frac{p^i}{p^0}_{x^i}f\frac{1}{p^0}\mathrm{\Gamma }_{\beta \gamma }^ip^\beta p^\gamma _{p^i}f=0$$ $$^\alpha _\alpha \varphi =0$$ $$G_{\alpha \beta }=8\pi T_{\alpha \beta }$$ $$T_{\alpha \beta }=_^3fp_\alpha p_\beta g^{\frac{1}{2}}\frac{dp^1dp^2dp^3}{p_0}+(_\alpha \varphi _\beta \varphi \frac{1}{2}g_{\alpha \beta }_\nu \varphi ^\nu \varphi )$$ where $`p_\alpha =g_{\alpha \beta }p^\beta `$, $`|g|`$ denotes the modulus of the determinant of the metric $`g_{\alpha \beta }`$, $`\mathrm{\Gamma }_{\alpha \beta }^\lambda `$ the Christoffel symbols, $`G_{\alpha \beta }`$ the Einstein tensor, and $`T_{\alpha \beta }`$ the energy-momentum tensor. Note that since the contribution of $`f`$ to the energy-momentum tensor is divergence-free , the form of the contribution of the scalar field to the energy-momentum tensor determines the field equation for $`\varphi `$. We refer to for the notion of spherical, plane and hyperbolic symmetry. We now consider a solution of the Einstein-Vlasov-scalar field system where all unknowns are invariant under one of these symmetries. We write the system in areal coordinates, i.e., coordinates are chosen such that $`R=t`$, where $`R`$ is the area radius function on a surface of symmetry. The circumstances under which coordinates of this type exist are discussed in . In such coordinates the metric $`g`$ takes the form $$ds^2=e^{2\mu (t,r)}dt^2+e^{2\lambda (t,r)}dr^2+t^2(d\theta ^2+\mathrm{sin}_k^2\theta d\phi ^2)$$ (1.1) where $$\mathrm{sin}_k\theta =\{\begin{array}{cc}\mathrm{sin}\theta \hfill & \text{for }k=1\text{ (spherical symmetry);}\hfill \\ 1\hfill & \text{for }k=0\text{ (plane symmetry);}\hfill \\ \mathrm{sinh}\theta \hfill & \text{for }k=1\text{ (hyperbolic symmetry)}\hfill \end{array}$$ $`t>0`$ denotes a time-like coordinate, $`r`$ and $`(\theta ,\phi )`$ range in the domains $`[0,\pi ]\times [0,2\pi ]`$, $`[0,2\pi ]\times [0,2\pi ]`$, $`[0,\mathrm{}[\times [0,2\pi ]`$ respectively, and stand for angular coordinates. The functions $`\lambda `$ and $`\mu `$ are periodic in $`r`$ with period $`1`$. It has been shown in that due to the symmetry, $`f`$ can be written as a function of $$t,r,w:=e^\lambda p^1\mathrm{and}F:=t^4[(p^2)^2+\mathrm{sin}_k^2\theta (p^3)^2],$$ i.e., $`f=f(t,r,w,F)`$. In these variables, we have $`p^0=e^\mu \sqrt{1+w^2+F/t^2}`$. The scalar field is a function of $`t`$ and $`r`$ which is periodic in $`r`$ with period 1. We denote by a dot and by a prime the derivatives of the metric components and of the scalar field with respect to $`t`$ and $`r`$ respectively. Using the results of , the complete Einstein-Vlasov-scalar field system can be written as follows: $$_tf+\frac{e^{\mu \lambda }w}{\sqrt{1+w^2+F/t^2}}_rf(\dot{\lambda }w+e^{\mu \lambda }\mu ^{}\sqrt{1+w^2+F/t^2})_wf=0$$ (1.2) $$e^{2\mu }(2t\dot{\lambda }+1)+k=8\pi t^2\rho $$ (1.3) $$e^{2\mu }(2t\dot{\mu }1)k=8\pi t^2p$$ (1.4) $$\mu ^{}=4\pi te^{\lambda +\mu }j$$ (1.5) $$e^{2\lambda }(\mu ^{\prime \prime }+\mu ^{}(\mu ^{}\lambda ^{}))e^{2\mu }(\ddot{\lambda }+(\dot{\lambda }+\frac{1}{t})(\dot{\lambda }\dot{\mu }))=4\pi q$$ (1.6) $$e^{2\lambda }\varphi ^{\prime \prime }e^{2\mu }\ddot{\varphi }e^{2\mu }(\dot{\lambda }\dot{\mu }+\frac{2}{t})\dot{\varphi }e^{2\lambda }(\lambda ^{}\mu ^{})\varphi ^{}=0$$ (1.7) where (1.7) is the wave equation in $`\varphi `$ and : $`\rho (t,r)=e^{2\mu }T_{00}(t,r)`$ $`={\displaystyle \frac{\pi }{t^2}}{\displaystyle _{\mathrm{}}^+\mathrm{}}{\displaystyle _0^+\mathrm{}}\sqrt{1+w^2+F/t^2}f(t,r,w,F)𝑑F𝑑w`$ (1.8) $`+{\displaystyle \frac{1}{2}}(e^{2\mu }\dot{\varphi }^2+e^{2\lambda }\varphi _{}^{}{}_{}{}^{2})`$ $`p(t,r)=e^{2\lambda }T_{11}(t,r)`$ $`={\displaystyle \frac{\pi }{t^2}}{\displaystyle _{\mathrm{}}^+\mathrm{}}{\displaystyle _0^+\mathrm{}}{\displaystyle \frac{w^2}{\sqrt{1+w^2+F/t^2}}}f(t,r,w,F)𝑑F𝑑w`$ (1.9) $`+{\displaystyle \frac{1}{2}}(e^{2\mu }\dot{\varphi }^2+e^{2\lambda }\varphi _{}^{}{}_{}{}^{2})`$ $$j(t,r)=e^{(\lambda +\mu )}T_{01}(t,r)=\frac{\pi }{t^2}_{\mathrm{}}^+\mathrm{}_0^+\mathrm{}wf(t,r,w,F)𝑑F𝑑we^{(\lambda +\mu )}\dot{\varphi }\varphi ^{}$$ (1.10) $`q(t,r)`$ $`={\displaystyle \frac{2}{t^2}}T_{22}(t,r)={\displaystyle \frac{2}{t^2\mathrm{sin}_k^2\theta }}T_{33}(t,r,\theta )`$ (1.11) $`={\displaystyle \frac{\pi }{t^4}}{\displaystyle _{\mathrm{}}^{\mathrm{}}}{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{F}{\sqrt{1+w^2+F/t^2}}}f(t,r,w,F)𝑑F𝑑w+e^{2\mu }\dot{\varphi }^2e^{2\lambda }\varphi _{}^{}{}_{}{}^{2}`$ We prescribe initial data at time $`t=1`$: $`f(1,r,w,F)=\stackrel{}{f}(r,w,F),\lambda (1,r)=\stackrel{}{\lambda }(r),\mu (1,r)=\stackrel{}{\mu }(r),`$ $`\varphi (1,r)=\stackrel{}{\varphi }(r),\dot{\varphi }(1,r)=\psi (r)`$ The choice $`t=1`$ is made only for convenience. Analogous results hold in the case of prescribed data on any hypersurface $`t=t_0>0`$. The paper is organized as follows. In section $`2`$, we show that the solution of the Cauchy problem corresponding to system (1.2)-(1.11) exists for all $`t]0,1]`$. In section $`3`$, we analyze the asymptotic behaviour of solutions as $`t0`$. The paper ends with a discussion of some interesting open problems. ## 2 Global existence in the past We use the continuation criterion in the following local existence result. ###### Theorem 2.1 Let $`\stackrel{}{f}C^1(^2\times [0,\mathrm{}[)`$ with $`\stackrel{}{f}(r+1,w,F)=\stackrel{}{f}(r,w,F)`$ for $`(r,w,F)^2\times [0,\mathrm{}[`$, $`\stackrel{}{f}0`$, and $`w_0:=sup\{|w||(r,w,F)\mathrm{supp}\stackrel{}{f}\}<\mathrm{}`$ $`F_0:=sup\{F|(r,w,F)\mathrm{supp}\stackrel{}{f}\}<\mathrm{}`$ Let $`\stackrel{}{\lambda },\psi C^1()`$, $`\stackrel{}{\mu },\stackrel{}{\varphi }C^2()`$ with $`\stackrel{}{\lambda }(r)=\stackrel{}{\lambda }(r+1)`$, $`\stackrel{}{\mu }(r)=\stackrel{}{\mu }(r+1)`$, $`\stackrel{}{\varphi }(r)=\stackrel{}{\varphi }(r+1)`$, $`\psi (r)=\psi (r+1)`$ and (1.5) satisfied for $`t=1`$. Then there exists a unique, left maximal, regular solution $`(f,\lambda ,\mu ,\varphi )`$ of system (1.2)-(1.11) with $`(f,\lambda ,\mu ,\varphi )(1)=(\stackrel{}{f},\stackrel{}{\lambda },\stackrel{}{\mu },\stackrel{}{\varphi })`$ and $`\dot{\varphi }(1)=\psi `$ on a time interval $`]T,1[`$ with $`T[0,1[`$. If 1. $`sup\{|w||(t,r,w,F)\mathrm{supp}f\}<\mathrm{}`$ , 2. $`sup\{(e^{2\mu }\dot{\varphi }^2+e^{2\lambda }\varphi ^2)(t,r);r\}<\mathrm{}`$, 3. $`\mu `$ is bounded, then $`T=0`$. If $`k0`$ or $`\stackrel{}{\mu }0`$ then condition $`3`$ is automatically satisfied. This is the content of theorems $`4.4`$ and $`4.5`$ in . For a regular solution, all derivatives which appear in the system exist and are continuous by definition (see ). In order to obtain the global existence of solutions, we prove the following results : ###### Lemma 2.2 Let $`D^+=e^\mu _t+e^\lambda _r`$ ; $`D^{}=e^\mu _te^\lambda _r`$; $`X=\dot{\varphi }e^\mu \varphi ^{}e^\lambda `$ ; $`Y=\dot{\varphi }e^\mu +\varphi ^{}e^\lambda `$ ; $`a=(\dot{\lambda }\frac{1}{t})e^\mu \mu ^{}e^\lambda `$ ; $`b=\frac{e^\mu }{t}`$ ; $`c=(\dot{\lambda }\frac{1}{t})e^\mu +\mu ^{}e^\lambda `$ and define $`X_2=e^\mu X`$ , $`Y_2=e^\mu Y`$. Then as a consequence of the field equations $`X`$ and $`Y`$ satisfy the system $$D^+X=aX+bY$$ (2.1) $$D^{}Y=bX+cY$$ (2.2) If in addition the field equations (1.3)-(1.4) are satisfied, then $`X_2`$ and $`Y_2`$ satisfy $$D^+X_2=e^\mu [\frac{k}{t}4\pi t(\rho p)]X_2\frac{e^\mu }{t}Y_2$$ (2.3) $$D^{}Y_2=\frac{e^\mu }{t}X_2+e^\mu [\frac{k}{t}4\pi t(\rho p)]Y_2$$ (2.4) Proof: This results from a straightforward calculation.$`\mathrm{}`$ ###### Lemma 2.3 Define $`X_2`$ and $`Y_2`$ as in lemma 2.2 and let $`B(t)`$ $`=sup\{(|X_2|^2+|Y_2|^2)^{1/2}(t,r);r\}`$ $`l(t)`$ $`=sup\{{\displaystyle \frac{1}{t}}+e^{2\mu }[{\displaystyle \frac{|k|}{t}}+4\pi t(\rho p)](t,r);r\}`$ If $`(X_2,Y_2)`$ is a solution of (2.3)-(2.4), then we obtain the estimate $$B(t)^2B(1)^2+2_t^1l(s)B(s)^2𝑑s$$ (2.5) with $`t`$ $``$ $`]T,1]`$, $`T>0`$. Proof: We deduce from system (2.3)-(2.4) : $$D^+X_2^2=2e^\mu \left[\frac{k}{t}4\pi t(\rho p)\right]X_2^22\frac{e^\mu }{t}X_2Y_2$$ $$D^{}Y_2^2=2\frac{e^\mu }{t}X_2Y_2+2e^\mu \left[\frac{k}{t}4\pi t(\rho p)\right]Y_2^2$$ On the corresponding characteristic curves $`(t,\gamma _i)`$, $`i=1,2`$ of the wave equation, (see ) $`D^+`$ or $`D^{}`$ is equal to $`e^\mu \frac{d}{dt}`$ and then $$\frac{d}{dt}X_2^2(t,\gamma _1(t))=2e^{2\mu }\left[\frac{k}{t}4\pi t(\rho p)\right]X_2^2(t,\gamma _1(t))\frac{2}{t}X_2Y_2(t,\gamma _1(t))$$ $$\frac{d}{dt}Y_2^2(t,\gamma _2(t))=\frac{2}{t}X_2Y_2(t,\gamma _2(t))+2e^{2\mu }\left[\frac{k}{t}4\pi t(\rho p)\right]Y_2^2(t,\gamma _2(t))$$ Integrate each of the two previous equations on $`[t,1]`$ and obtain respectively : $`X_2^2(t,\gamma _1(t))`$ $`=X_2^2(1,\gamma _1(1))+2{\displaystyle _t^1}\{e^{2\mu }[{\displaystyle \frac{k}{s}}+4\pi s(\rho p)]X_2^2+{\displaystyle \frac{1}{s}}X_2Y_2\}(s,\gamma _1(s))𝑑s`$ $`X_2^2(1,\gamma _1(1))+2{\displaystyle _t^1}\{[{\displaystyle \frac{1}{2s}}+e^{2\mu }({\displaystyle \frac{k}{s}}+4\pi s(\rho p))]X_2^2+{\displaystyle \frac{1}{2s}}Y_2^2\}(s,\gamma _1(s))𝑑s;`$ $$Y_2^2(t,\gamma _2(t))Y_2^2(1,\gamma _2(1))+2_t^1\{\frac{1}{2s}X_2^2+\left[\frac{1}{2s}+e^{2\mu }(\frac{k}{s}+4\pi s(\rho p))\right]Y_2^2\}(s,\gamma _2(s))𝑑s$$ Add the two previous inequalities and take the supremum over space to obtain estimate (2.5).$`\mathrm{}`$ Unless otherwise specified in what follows constants denoted by $`C`$ will be positive, may depend on the initial data and may change their value from line to line. ###### Proposition 2.4 Let $`(f,\lambda ,\mu ,\varphi )`$ be a solution of the full system (1.2)-(1.11) on a left maximal interval of existence $`]T,1]`$, $`T>0`$, with initial data as in theorem 2.1. If 1. $`Q(t)=sup\{|w||(r,w,f)\mathrm{supp}f(t),t]T,1]\}<\mathrm{}`$ 2. $`\mu `$ is bounded. then $`T=0`$. If $`k0`$ or $`\stackrel{}{\mu }0`$ then condition $`2`$ is automatically satisfied. Proof: We need to prove that $`K(t)=sup\{(|X|^2+|Y|^2)^{1/2}(t,r);r\}`$ is bounded for all $`t]T,1]`$; where $`X`$ and $`Y`$ are defined in lemma 2.2. Subtract the two equations (1.8)-(1.9) to obtain : $`\rho p`$ $`={\displaystyle \frac{\pi }{t^2}}{\displaystyle _{\mathrm{}}^{\mathrm{}}}{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{1+F/t^2}{\sqrt{1+w^2+F/t^2}}}f𝑑F𝑑w`$ (2.6) $`{\displaystyle \frac{\pi }{t^2}}{\displaystyle _{\mathrm{}}^{\mathrm{}}}{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{1+F/t^2}{\sqrt{1+F/t^2}}}f𝑑F𝑑w`$ $`{\displaystyle \frac{\pi }{t^2}}{\displaystyle _{\mathrm{}}^{\mathrm{}}}{\displaystyle _0^{\mathrm{}}}\sqrt{1+F/t^2}f𝑑F𝑑w`$ $`{\displaystyle \frac{C}{t^3}};\text{since}Q(t)<\mathrm{}`$ Using $`(4.15)`$ of , which shows that $`e^{2\mu }=O(t)`$, we obtain : $$l(t)\frac{C}{t}+\frac{|k|}{C}$$ Then (2.5) implies : $$B(t)^2B(1)^2+2_t^1(C(1+\frac{1}{s})+\frac{1}{s})B(s)^2𝑑s$$ And by Gronwall’s lemma, $$B(t)^2Ct^{C2}$$ We have from (1.9), $`p(s,r)`$ $`{\displaystyle \frac{\pi }{s^2}}{\displaystyle _{\mathrm{}}^+\mathrm{}}{\displaystyle _0^+\mathrm{}}{\displaystyle \frac{w^2}{|w|}}f(s,r,w,F)𝑑F𝑑w+{\displaystyle \frac{1}{4}}e^{2\mu }(X_2^2+Y_2^2)(s,r)`$ $`{\displaystyle \frac{C}{s^2}}+{\displaystyle \frac{1}{4}}e^{2\mu }B(s)^2`$ $`Cs^2+Cs^{C2}e^{2\mu }`$ Then using (1.4), we obtain the estimate : $`e^{2\mu (t,r)}`$ $`={\displaystyle \frac{e^{2\stackrel{}{\mu }(r)}+k}{t}}k{\displaystyle \frac{8\pi }{t}}{\displaystyle _1^t}s^2p(s,r)𝑑s`$ $`{\displaystyle \frac{e^{2\stackrel{}{\mu }(r)}+|k|}{t}}+{\displaystyle \frac{C}{t}}{\displaystyle _t^1}(1+s^Ce^{2\mu })𝑑s`$ $`Ct^1(1+{\displaystyle _t^1}s^Ce^{2\mu }𝑑s).`$ By Gronwall’s lemma, we deduce that $`e^{2\mu }`$ $`Ct^1\mathrm{exp}[\frac{C}{1C}(1t^{1C})]`$. Therefore, $`(X^2+Y^2)(t,r)=e^{2\mu }(X_2^2+Y_2^2)(t,r)`$ $`Ct^{3C}\mathrm{exp}[\frac{C}{1C}(1t^{1C})]`$ i.e $`K(t)`$ is bounded. We conclude by theorem 2.1 that $`T=0`$.$`\mathrm{}`$ In the next theorem, we prove that the solution exists on the full interval $`]0,1]`$. ###### Theorem 2.5 Consider a solution of the Einstein-Vlasov-scalar field system with $`k0`$ and initial data given for $`t=1`$. Then this solution exists on the whole interval $`]0,1]`$. If $`k<0`$ and $`\stackrel{}{\mu }0`$, the same result holds. Proof : The strategy of the proof is the following : suppose we have a solution on an interval $`]T,1]`$ with $`T>0`$. We want to show that the solution can be extended to the past. By consideration of the maximal interval of existence this will prove the assertion. Firstly let us prove that under the hypotheses of the theorem, $`\mu `$ is bounded above. From the field equation (1.4) and since $`p(s,r)0`$, we have for $`k0`$, $`e^{2\mu (t,r)}{\displaystyle \frac{e^{2\stackrel{}{\mu }(r)}+k}{t}}ke^{2\stackrel{}{\mu }(r)}`$ For the case $`k=1`$, $`e^{2\mu }\frac{e^{2\stackrel{}{\mu }}1}{t}+11`$ which gives the upper bound of $`\mu `$ for $`\stackrel{}{\mu }0`$. In either case, $`\mu `$ is bounded above. Now, let us prove that $`w`$ is bounded. Consider the following rescaled version of $`w`$, called $`u_1`$, which has been inspired by the works of (p. 1090) and (p. 5336): $$u_1=\frac{e^\mu }{2t}w.$$ If we prove that $`\mu `$ is bounded below then the boundedness of $`u_1`$ will imply the boundedness of $`w`$. So let us show that $`\mu `$ is bounded below under the assumption that $`u_1`$ is bounded. We have $$\frac{d}{dt}(te^{2\mu })=k8\pi t^2p.$$ (2.7) Transforming the integral term defining $`p`$ to $`u_1`$ as an integration variable instead of $`w`$ yields $$p=_{\mathrm{}}^{\mathrm{}}_0^{\mathrm{}}\frac{8\pi te^{3\mu }u_1^2}{\sqrt{1+4t^2e^{2\mu }u_1^2+F/t^2}}f𝑑F𝑑u_1+\frac{1}{4}(X^2+Y^2);$$ where $`X`$ and $`Y`$ are defined in lemma 2.2. The integrand in the first term in $`p`$ can then be estimated by $`4\pi e^{2\mu }|u_1|`$. We have from (1.8)-(1.9): $`e^{2\mu }(\rho p)`$ $`={\displaystyle \frac{\pi }{t^2}}{\displaystyle _{\mathrm{}}^{\mathrm{}}}{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{(1+F/t^2)2te^\mu }{\sqrt{1+4t^2e^{2\mu }u_1^2+F/t^2}}}f𝑑F𝑑u_1`$ $`{\displaystyle \frac{2\pi }{t}}{\displaystyle _{\mathrm{}}^{\mathrm{}}}{\displaystyle _0^{\mathrm{}}}(1+F/t^2)e^\mu f𝑑F𝑑u_1`$ $`Ct^3\overline{u}_1e^\mu `$ where $`\overline{u}_1`$ is the maximum modulus of $`u_1`$ on the support of $`f`$ at a given time. We can then estimate from (2.5) $`l(s)`$ by $`h(s)`$ $`=`$ $`Csup\{1+s^1+s^2e^\mu \overline{u}_1(s,r);r\}`$ and $`B(t)^2`$ by $`B(1)^2\mathrm{exp}(_t^1h(s)𝑑s)`$. Thus $$X^2+Y^2e^{2\mu }B(t)^2e^{2\mu }B(1)^2\mathrm{exp}\left(_t^1h(s)𝑑s\right).$$ Therefore, using the bound for $`\mu `$ and $`u_1`$, $`p`$ can be estimated by $`Ce^{2\mu }`$ and so (2.7) implies that $$|\frac{d}{dt}(te^{2\mu })|C(1+te^{2\mu }),$$ Integrating this with respect to $`t`$ over $`[t,1]`$ and using the Gronwall inequality implies that $`te^{2\mu }`$ is bounded on $`]T,1]`$; that is $`\mu `$ is bounded below on the given time interval. The next step is to prove that $`u_1`$ is bounded. To this end, it suffices to get a suitable integral inequality for $`\overline{u}_1`$. Since $`u_1=u_1(t,r(t))`$, we can compute $`\dot{u}_1`$ : $$\dot{u}_1=\frac{e^\mu }{2t^2}w+\frac{e^\mu }{2t}w(\dot{\mu }+\dot{r}\mu ^{})+\frac{e^\mu }{2t}\dot{w}$$ i.e. $$\dot{u}_1=\left(\dot{\mu }+\dot{r}\mu ^{}\frac{1}{t}\right)u_1+\frac{e^\mu }{2t}\dot{w}$$ (2.8) We have $$\mu ^{}=4\pi te^{\mu +\lambda }j,\dot{r}=\frac{e^{\mu \lambda }w}{\sqrt{1+w^2+F/t^2}}$$ and $$\dot{w}=4\pi te^{2\mu }(j\sqrt{1+w^2+F/t^2}\rho w)+\frac{1+ke^{2\mu }}{2t}w$$ so that (2.8) implies the following : $$\dot{u_1}|u_1|=e^{2\mu }\left[4\pi t(\rho p)+\frac{k}{t}\right]u_1|u_1|+2\pi e^{3\mu }j\frac{(1+F/t^2)|u_1|}{\sqrt{1+4t^2e^{2\mu }u_1^2+F/t^2}}$$ (2.9) In order to estimate the modulus of the first term on the right hand side of equation (2.9), we need an estimate of $`e^{2\mu }(\rho p)\overline{u}_1^2`$. For convenience let $`\mathrm{log}_+`$ be defined by $`\mathrm{log}_+(x)=\mathrm{log}x`$ when $`\mathrm{log}x`$ is positive and $`\mathrm{log}_+(x)=0`$, otherwise. Then estimating the integral defining $`\rho p`$ shows that $$\rho pC(1+\mathrm{log}_+(\overline{w})),$$ i.e. $$\rho pC(1+\mathrm{log}_+(\overline{u}_1)\mu ).$$ The expression $`\mu `$ is not under control ; however the expression we wish to estimate contains a factor $`e^{2\mu }`$. The function $`\mu \mu e^{2\mu }`$ has an absolute maximum at $`1/2`$ which is $`(1/2)e^1`$. Thus the first term on the right hand side of equation (2.9) can be estimated by $`C\overline{u}_1^2(1+\mathrm{log}_+(\overline{u}_1))`$. Next the second term on the right hand side of equation (2.9) will be estimated. By definition $$j=\frac{\pi }{t^2}_{\mathrm{}}^{\mathrm{}}_0^{\mathrm{}}wf(t,r,w,F)𝑑F𝑑w\dot{\varphi }\varphi ^{}e^{\mu \lambda }=j_1+j_2$$ The first term of $`j`$ can be estimated by $`C\overline{w}^2`$, i.e. $$|j_1|C\overline{u}_1^2e^{2\mu }$$ and the second term $`|j_2|`$ $`\frac{1}{2}e^{2\mu }B(t)^2`$; so that it suffices to estimate the quantity $$\frac{(\overline{u}_1^2+B(t)^2)(1+F/t^2)}{\sqrt{1+4t^2e^{2\mu }u_1^2+F/t^2}}|u_1|$$ (2.10) in order to estimate the second term on the right hand side of equation (2.9). But since $`\mu `$ and $`t^1`$ are bounded on the interval being considered, the quantity (2.10) can be estimated by $`C(\overline{u}_1^2+B(t)^2)`$. Thus adding the estimates for the first and second terms on the right hand side of (2.9) allows us to deduce from (2.9) that $$|\dot{u}_1||u_1|C\overline{u}_1^2(1+\mathrm{log}_+(\overline{u}_1))+C(\overline{u}_1^2+B(t)^2)$$ i.e $$|\frac{d}{dt}|u_1|^2|C(\overline{u}_1)^2(1+\mathrm{log}_+(\overline{u}_1^2))+CB(t)^2$$ Integrating over $`[t,1]`$ gives : $$\overline{u}_1^2(t)\overline{u}_1^2(1)+C_t^1\left[\overline{u}_1^2(s)(1+\mathrm{log}_+(\overline{u}_1^2(s)))+B(s)^2\right]𝑑s$$ (2.11) We deduce from the estimate of $`\rho p`$ and from inequality (2.5), that $`l(s)`$ can be estimated by $`C(1+\mathrm{log}_+(\overline{u}_1))`$. We then obtain $$B(t)^2B(1)^2+C_t^1(1+\mathrm{log}_+(\overline{u}_1^2)(s))B(s)^2𝑑s$$ (2.12) Adding (2.11) and (2.12) gives estimate : $$\overline{u}_1^2(t)+B(t)^2\overline{u}_1^2(1)+B(1)^2+C_t^1(1+\overline{u}_1^2+B(s)^2)\left[1+\mathrm{log}_+(1+\overline{u}_1^2(s)+B(s)^2)\right]𝑑s$$ Set $`v(t)=\overline{u}_1^2(t)+B(t)^2`$, then the above estimate can be written $$v(t)v(1)+C_t^1(1+v(s))\left[1+\mathrm{log}_+(1+v(s))\right]𝑑s$$ (2.13) By the comparison principle for solutions of integral equations, it is enough to show that the solution of the integral equation $$a(t)=a(1)+C_t^1(1+a(s))\left(1+\mathrm{log}_+(1+a(s))\right)𝑑s$$ is bounded. The solution $`a(t)`$ is a non-increasing function. Thus either $`a(t)e`$ everywhere, in which case the desired conclusion is immediate or there is some $`T_1`$ in $`]T,1]`$ such that $`ea(t)`$ on $`]T,T_1]`$. We take $`T_1`$ maximal with that property. Then it follows on $`]T,T_1]`$ that inequality $$a(t)C\left(1+_t^{T_1}a(s)(1+\mathrm{log}a(s))𝑑s\right)$$ holds for a constant C. The boundedness of $`a(t)`$ follows from that of the solution of the differential equation $`\dot{x}=Cx(1+\mathrm{log}x)`$ which is exp(exp($`Ct`$)$`1`$). In either case $`a(t)`$ is bounded. Thus $`\overline{u}_1^2`$ and $`B(t)^2`$ are bounded i.e $`w`$ and $`K(t)`$ are bounded. The proof of the theorem is complete using proposition 2.4.$`\mathrm{}`$ ## 3 On past asymptotic behaviour In this section we examine the behaviour of solutions as $`t0`$. Firstly we generalize the work of Ringström (P. S310-S311) to bound the quantity $`|\varphi ^{}|e^{\mu \lambda }`$ by $`C|t\mathrm{log}t|^1`$, where $`C`$ is a positive constant. ###### Lemma 3.1 Let $`A_1=\frac{1}{8}(\dot{\varphi }+\frac{\varphi }{t\mathrm{log}t}+\varphi ^{}e^{\mu \lambda })^2`$ and $`A_2=\frac{1}{8}(\dot{\varphi }+\frac{\varphi }{t\mathrm{log}t}\varphi ^{}e^{\mu \lambda })^2`$ with $`t]0,1[`$. If $`\varphi `$ satisfies the wave equation, then $`(_t+e^{\mu \lambda }_r)A_1={\displaystyle \frac{1}{4t}}(1+{\displaystyle \frac{1}{\mathrm{log}t}})[(\dot{\varphi }+{\displaystyle \frac{\varphi }{t\mathrm{log}t}})^2+\varphi ^2e^{2\mu 2\lambda }]`$ (3.1) $`+{\displaystyle \frac{1}{2t}}(1+{\displaystyle \frac{1}{\mathrm{log}t}})\varphi ^2e^{2\mu 2\lambda }+{\displaystyle \frac{1}{4}}(\dot{\lambda }\dot{\mu }+{\displaystyle \frac{1}{t}})(\dot{\varphi }\varphi ^{}e^{\mu \lambda })(\dot{\varphi }+{\displaystyle \frac{\varphi }{t\mathrm{log}t}}+\varphi ^{}e^{\mu \lambda })`$ $`(_t+e^{\mu \lambda }_r)A_2={\displaystyle \frac{1}{4t}}(1+{\displaystyle \frac{1}{\mathrm{log}t}})[(\dot{\varphi }+{\displaystyle \frac{\varphi }{t\mathrm{log}t}})^2+\varphi ^2e^{2\mu 2\lambda }]`$ (3.2) $`+{\displaystyle \frac{1}{2t}}(1+{\displaystyle \frac{1}{\mathrm{log}t}})\varphi ^2e^{2\mu 2\lambda }+{\displaystyle \frac{1}{4}}(\dot{\lambda }\dot{\mu }+{\displaystyle \frac{1}{t}})(\dot{\varphi }+\varphi ^{}e^{\mu \lambda })(\dot{\varphi }+{\displaystyle \frac{\varphi }{t\mathrm{log}t}}\varphi ^{}e^{\mu \lambda })`$ Proof : This results from a straightforward calculation.$`\mathrm{}`$ ###### Proposition 3.2 Let $`(f,\lambda ,\mu ,\varphi )`$ be a left maximal solution of the Einstein-Vlasov-scalar field system on the interval $`]T,1]`$, $`0T<e^1`$. Assume that $`Q(t)=sup\{w|(r,w,F)\mathrm{supp}f(t)\}Ct^\alpha `$ for some positive constants $`C`$, $`\alpha `$ and for some $`t]T,e^1]`$. Then $$(\dot{\varphi }+\frac{\varphi }{t\mathrm{log}t})^2+\varphi ^2e^{2\mu 2\lambda }C(t\mathrm{log}t)^2$$ (3.3) Proof : Consider the two characteristic curves $`(t,\gamma _1(t))`$ and $`(t,\gamma _2(t))`$ of the wave operator. Since $`t]0,e^1]`$, the term $`(1+\frac{1}{\mathrm{log}t})\varphi ^2e^{2\mu 2\lambda }`$ is nonnegative and $`(\dot{\varphi }+\frac{\varphi }{t\mathrm{log}t})^2+\varphi ^2e^{2\mu 2\lambda }=4(A_1+A_2)`$, then from (3.1) : $`(_t`$ $`+e^{\mu \lambda }_r)A_1(t,\gamma _1(t)){\displaystyle \frac{1}{t}}(1+{\displaystyle \frac{1}{\mathrm{log}t}})(A_1+A_2)(t,\gamma _1(t))`$ (3.4) $`{\displaystyle \frac{1}{4}}(\dot{\lambda }\dot{\mu }+{\displaystyle \frac{1}{t}})(\dot{\varphi }+\varphi ^{}e^{\mu \lambda })(\dot{\varphi }+{\displaystyle \frac{\varphi }{t\mathrm{log}t}}+\varphi ^{}e^{\mu \lambda })(t,\gamma _1(t))`$ Similarly, we deduce from (3.2) that : $`(_t`$ $`e^{\mu \lambda }_r)A_2(t,\gamma _2(t)){\displaystyle \frac{1}{t}}(1+{\displaystyle \frac{1}{\mathrm{log}t}})(A_1+A_2)(t,\gamma _2(t))`$ (3.5) $`{\displaystyle \frac{1}{4}}(\dot{\lambda }\dot{\mu }+{\displaystyle \frac{1}{t}})(\dot{\varphi }\varphi ^{}e^{\mu \lambda })(\dot{\varphi }+{\displaystyle \frac{\varphi }{t\mathrm{log}t}}\varphi ^{}e^{\mu \lambda })(t,\gamma _2(t))`$ Since $`Q(t)Ct^\alpha `$, we can bound $`\rho p`$ by $`Ct^{3+\alpha }`$ (see (2.6)). $`e^{2\mu }Ct`$; then subtracting (1.3) -(1.4) gives $`(\dot{\lambda }\dot{\mu })(t)+\frac{1}{t}`$ $`=\frac{ke^{2\mu }}{t}+4\pi te^{2\mu }(\rho p)`$ $`C(1+t^{1+\alpha })`$; and from (2.5), $`l(s)`$ can be bounded by $`s^1+C+Cs^{1+\alpha }`$. We deduce from (2.5) (consider the integral term in the interval $`[t,e^1]`$) that $`B(t)^2B(e^1)^2\mathrm{exp}[2_t^{e^1}(s^1+C+Cs^{1+\alpha })𝑑s]`$ i.e., $`B(t)^2Ct^2`$. Therefore $`|\dot{\varphi }(t)|`$ and $`|\varphi ^{}|e^{\mu \lambda }(t)`$ are bounded each by $`Ct^1`$. We can then have a lower bound of the second term of the right hand side of each inequality (3.4) and (3.5) which is $`C(t^2+t^{3+\alpha })`$. Then $$(_t+e^{\mu \lambda }_r)A_1(t,\gamma _1(t))\frac{1}{t}(1+\frac{1}{\mathrm{log}t})(A_1+A_2)(t,\gamma _1(t))C(t^2+t^{3+\alpha })$$ and $$(_te^{\mu \lambda }_r)A_2(t,\gamma _2(t))\frac{1}{t}(1+\frac{1}{\mathrm{log}t})(A_1+A_2)(t,\gamma _2(t))C(t^2+t^{3+\alpha })$$ On the corresponding characteristic, we have $`_t+e^{\mu \lambda }_r`$ or $`_te^{\mu \lambda }_r`$ equal to $`\frac{d}{dt}`$. Take the supremum in the space of each of the above two inequalities and add them. Then $$\frac{d}{dt}(A_1+A_2)(t,r)\frac{2}{t}(1+\frac{1}{\mathrm{log}t})(A_1+A_2)(t,r)C(t^2+t^{3+\alpha })$$ Set $`u(t)=(A_1+A_2)(t)`$ and $`v(t)=(t\mathrm{log}t)^2u(t)`$. If $`v(t)`$ is bounded, then we conclude that $`u(t)`$ is bounded by $`C(t\mathrm{log}t)^2`$. Let us prove that $`v(t)`$ is bounded. We have : $`{\displaystyle \frac{dv}{dt}}`$ $`=(t\mathrm{log}t)^2({\displaystyle \frac{du}{dt}}+{\displaystyle \frac{2}{t}}(1+{\displaystyle \frac{1}{\mathrm{log}t}})u)`$ $`C(\mathrm{log}t)^2(1+t^{1+\alpha })`$ Then, $`v(t)`$ $`v(e^1)+C_t^{e^1}(1+s^{1+\alpha })(\mathrm{log}s)^2𝑑s`$ $`v(e^1)+C`$. We obtain the desired conclusion of the proposition.$`\mathrm{}`$ In the case $`f=0`$, we obtain from the previous proposition, theorems 2.5 and 2.1, the global existence of solutions and the above estimates hold on the whole interval $`]0,e^1]`$. In general we do not know how to use this proposition to obtain precise asymptotics. It seems that if the estimates could be improved slightly they would allow a bootstrap argument on the bound for $`Q`$ similar to that used in . These estimates have been included here in the hope that they might help someone else to complete the argument. Next we prove that the curvature invariant $`R_{\alpha \beta \gamma \delta }R^{\alpha \beta \gamma \delta }`$ called the Kretschman scalar blows up as $`t0`$. Then there is a spacetime singularity and the spacetime cannot be extended further. ###### Theorem 3.3 Let $`(f,\lambda ,\mu ,\varphi )`$ be a regular solution of the surface-symmetric Einstein-Vlasov-scalar field system on the interval $`]0,1]`$ with data given for $`t=1`$. Then $$(R_{\alpha \beta \gamma \delta }R^{\alpha \beta \gamma \delta })(t,r)\frac{4}{t^6}\left(infe^{2\stackrel{}{\mu }}+k\right)^2,$$ (3.6) with $`r`$. Proof We can use the following expression for the Kretschman scalar from . $`R_{\alpha \beta \gamma \delta }R^{\alpha \beta \gamma \delta }`$ $`=4[e^{2\lambda }(\mu ^{\prime \prime }+\mu ^{}(\mu ^{}\lambda ^{}))e^{2\mu }(\ddot{\lambda }+\dot{\lambda }(\dot{\lambda }\dot{\mu }))]^2`$ $`+{\displaystyle \frac{8}{t^2}}[e^{4\mu }\dot{\lambda }^2+e^{4\mu }\dot{\mu }^22e^{2(\lambda +\mu )}(\mu ^{})^2]`$ $`+{\displaystyle \frac{4}{t^4}}(e^{2\mu }+k)^2`$ $`=:K_1+K_2+K_3`$ The first term $`K_1`$ is nonnegative and can be dropped. Inserting the expressions $$e^{2\mu }\dot{\lambda }=4\pi t\rho \frac{k+e^{2\mu }}{2t};e^{2\mu }\dot{\mu }=4\pi tp+\frac{k+e^{2\mu }}{2t};e^{\lambda \mu }\mu ^{}=4\pi tj$$ into the formula for $`K_2`$ yields $`K_2={\displaystyle \frac{8}{t^2}}\left[16\pi ^2t^2(\rho ^2+p^22j^2)4\pi t(\rho p){\displaystyle \frac{k+e^{2\mu }}{t}}+{\displaystyle \frac{(k+e^{2\mu })^2}{2t^2}}\right].`$ Now $`|j(t,r)|`$ $`{\displaystyle \frac{\pi }{t^2}}{\displaystyle _{\mathrm{}}^{\mathrm{}}}{\displaystyle _0^{\mathrm{}}}(1+w^2+F/t^2)^{1/4}f^{1/2}{\displaystyle \frac{|w|}{(1+w^2+F/t^2)^{1/4}}}f^{1/2}𝑑F𝑑w`$ $`+{\displaystyle \frac{1}{2}}(\dot{\varphi }^2e^{2\mu }+\varphi ^2e^{2\lambda })`$ $`{\displaystyle \frac{\pi }{t^2}}[{\displaystyle _{\mathrm{}}^{\mathrm{}}}{\displaystyle _0^{\mathrm{}}}\sqrt{1+w^2+F/t^2}f𝑑F𝑑w]^{\frac{1}{2}}[{\displaystyle _{\mathrm{}}^{\mathrm{}}}{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{w^2}{\sqrt{1+w^2+F/t^2}}}f𝑑F𝑑w]^{\frac{1}{2}}`$ $`+{\displaystyle \frac{1}{2}}(\dot{\varphi }^2e^{2\mu }+\varphi ^2e^{2\lambda })\text{by the Cauchy-Schwarz inequality.}`$ $`{\displaystyle \frac{1}{2}}(\rho +p)(t,r).`$ In fact the above inequality holds in general for all choices of matter satisfying the dominant energy condition. Therefore $`K_2`$ $`{\displaystyle \frac{8}{t^2}}\left[8\pi t^2(\rho p)^24\pi t(\rho p){\displaystyle \frac{k+e^{2\mu }}{t}}+{\displaystyle \frac{(k+e^{2\mu })^2}{2t^2}}\right]`$ $`{\displaystyle \frac{4}{t^2}}[4\pi t(\rho p){\displaystyle \frac{k+e^{2\mu }}{t}}]^20.`$ Recalling the expression for $`e^{2\mu }`$ we get $`e^{2\mu }+k`$ $`={\displaystyle \frac{e^{2\stackrel{}{\mu }(r)}+k}{t}}+{\displaystyle \frac{8\pi }{t}}{\displaystyle _t^1}s^2p(s,r)𝑑s`$ $`{\displaystyle \frac{e^{2\stackrel{}{\mu }}+k}{t}}{\displaystyle \frac{infe^{2\stackrel{}{\mu }}+k}{t}}`$ and thus $$K_3=\frac{4}{t^4}(e^{2\mu }+k)^2\frac{4}{t^6}\left(infe^{2\stackrel{}{\mu }}+k\right)^2$$ We obtain (3.6) and deduce that $$\underset{t0}{lim}(R_{\alpha \beta \gamma \delta }R^{\alpha \beta \gamma \delta })(t,r)=\mathrm{},$$ uniformly in $`r`$. Next we prove that the singularity at $`t=0`$ is a crushing singularity i.e. the mean curvature of the surfaces of constant $`t`$ blows up. In the case where there is only a scalar field and no Vlasov contribution this singularity is velocity dominated i.e the generalized Kasner exponents have limits as $`t0`$. ###### Theorem 3.4 Let $`(f,\lambda ,\mu ,\varphi )`$ be a regular solution of the surface-symmetric Einstein-Vlasov-scalar field system on the interval $`]0,1]`$ with initial data given on $`t=1`$. Let $$K(t,r):=e^\mu \left(\dot{\lambda }(t,r)+\frac{2}{t}\right)$$ denote the mean curvature of the hypersurfaces of constant $`t`$. Then $$K(t,r)Ct^{3/2},$$ where $`C`$ is a positive constant. Proof We use the same argument as in and obtain the following : $`K(t,r)=(\dot{\lambda }+\frac{2}{t})e^\mu `$; $`\dot{\lambda }=e^{2\mu }\left(4\pi t\rho \frac{k+e^{2\mu }}{2t}\right)e^{2\mu }\left(\frac{k+e^{2\mu }}{2t}\right)`$ and $$K(t,r)\frac{k3e^{2\mu }}{2t}e^\mu .$$ For $`k=0`$ or $`k=1`$, $$K(t,r)\frac{3}{2t}e^\mu .$$ and the estimate $`e^{2\mu }{\displaystyle \frac{e^{2\stackrel{}{\mu }}+k}{t}}`$ implies $`K(t,r){\displaystyle \frac{3}{2t}}({\displaystyle \frac{infe^{2\stackrel{}{\mu }}+k}{t}})^{1/2}Ct^{3/2}\text{where }C=\frac{3}{2}(infe^{2\stackrel{}{\mu }}+k)^{1/2}.`$ For $`k=1`$ we have $`e^{2\mu }{\displaystyle \frac{e^{2\stackrel{}{\mu }}}{t}}>1=k`$ thus $`K(t,r)`$ $`({\displaystyle \frac{e^{2\mu }3}{2}}){\displaystyle \frac{e^\mu }{t}}`$ $`{\displaystyle \frac{e^\mu }{t}}{\displaystyle \frac{infe^\stackrel{}{\mu }}{t^{3/2}}}`$ $`Ct^{3/2}\text{where }C=infe^\stackrel{}{\mu }.`$ We deduce from above that $$\underset{t0}{lim}K(t,r)=\mathrm{},$$ uniformly in $`r`$. ###### Theorem 3.5 Let $`(\lambda ,\mu ,\varphi )`$ be a regular solution of the Einstein-scalar field system with spherical, plane or hyperbolic symmetry on the interval $`]0,1]`$ with initial data given at $`t=1`$. Then $$\underset{t0}{lim}\frac{K_1^1(t,r)}{K(t,r)}=a(r);\underset{t0}{lim}\frac{K_2^2(t,r)}{K(t,r)}=\underset{t0}{lim}\frac{K_3^3(t,r)}{K(t,r)}=\frac{1}{2}(1a(r)),$$ uniformly in $`r`$, where $$\frac{K_1^1(t,r)}{K(t,r)},\frac{K_2^2(t,r)}{K(t,r)},\frac{K_3^3(t,r)}{K(t,r)}$$ are the generalized Kasner exponents and $`a(r)`$ a continuous function of $`r`$. Proof We have as in $$\frac{K_1^1(t,r)}{K(t,r)}=\frac{t\dot{\lambda }(t,r)}{t\dot{\lambda }(t,r)+2};\frac{K_2^2(t,r)}{K(t,r)}=\frac{K_3^3(t,r)}{K(t,r)}=\frac{1}{t\dot{\lambda }(t,r)+2}.$$ As we have seen previously $$e^{2\mu (t,r)}Ct$$ which implies that $$e^{2\mu (t,r)}0\text{as }t0$$ Let $`t_0]0,e^1]`$ and $`t]0,t_0]`$. From (3.3), $$_t(\frac{\varphi }{\mathrm{log}t})=\frac{1}{\mathrm{log}t}(\dot{\varphi }\frac{\varphi }{t\mathrm{log}t})=O(t^1(\mathrm{log}t)^2)$$ Since $`t^1(\mathrm{log}t)^2`$ is integrable on the interval $`(0,t_0]`$ it follows that we can define $$A(r)=\underset{t0}{lim}\frac{\varphi (t,r)}{\mathrm{log}t}=\frac{\varphi (t_0,r)}{\mathrm{log}t_0}_0^{t_0}(\mathrm{log}s)^1(\dot{\varphi }(s,r)\varphi (s,r)/s\mathrm{log}s)𝑑s$$ Since from (3.3), $`(\dot{\varphi }\frac{\varphi }{t\mathrm{log}t})=O((t|\mathrm{log}t|)^1)`$, we have $$t\dot{\varphi }=\frac{\varphi }{\mathrm{log}t}+O((|\mathrm{log}t|)^1)$$ so that $`t\dot{\varphi }A(r)`$ as $`t0`$. Inequality (3.3) shows also that $`\varphi ^2e^{2\mu 2\lambda }`$ $`=O((t|\mathrm{log}t|)^2)`$. Using these limits, we have $`t\dot{\lambda }(t,r)=2\pi (t^2\dot{\varphi }^2+t^2\varphi ^2e^{2\mu 2\lambda }){\displaystyle \frac{k}{2}}e^{2\mu }{\displaystyle \frac{1}{2}}2\pi A(r)^2{\displaystyle \frac{1}{2}}\text{as }t0\text{, uniformly in }r.`$ We take $`a(r)=\frac{4\pi A(r)^21}{4\pi A(r)^2+3}`$ to complete the proof. $`\mathrm{}`$ ## 4 Discussion and outlook It is an open problem to remove the restriction on $`\mu `$ in Theorem 2.5, even when the scalar field vanishes. The only example where existence up to $`t=0`$ is known to fail is a vacuum solution, the pseudo-Schwarzschild solution (cf. the discussion in , p. 115). Perhaps it can only fail in the vacuum case. As mentioned in the introduction, the result for the plane symmetric case extends to solutions of the Einstein-Vlasov system with $`T^2`$-symmetry . Once existence up to $`t=0`$ is known the ideal goal is to obtain detailed information on the asymptotics. For the Einstein-Vlasov system this has been done in for small data. In Proposition 3.2 an analogue of some parts of the proof of the small data theorem in is obtained but this is not sufficient in order to determine the asymptotic behaviour. There is a formal similarity between the Einstein-Vlasov-scalar field system with plane symmetry and the Einstein-Vlasov system with polarized Gowdy symmetry. If the problem of asymptotics could be solved for the first problem it would probably lead to valuable insights for the second problem. In the case of general Gowdy symmetry, results on asymptotics are available in the vacuum case but they are hard to obtain . In the yet more general case of $`T^2`$-symmetry the only thing known is a construction of a class of vacuum solutions with prescribed asymptotics . Acknowledgements : The authors acknowledge support by a research grant from the VolkswagenStiftung, Federal Republic of Germany. The major part of this work was completed during a visit of D. Tegankong to the Max-Planck Institute for Gravitational Physics, Golm, Germany.
warning/0506/astro-ph0506465.html
ar5iv
text
# STEREO ARRAY of 30 m Imaging Atmospheric Čerenkov Telescopes: A Next-Generation Detector for Ground-Based High Energy Gamma-ray Astronomy ## 1 Introduction Development of further instrumentation in the field of very high energy (VHE) $`\gamma `$-ray astronomy is primarily motivated by the physics goals as perceived by the astrophysics community today (Weekes 2004). Among those one has to mention (i) observations of the supernova remnants (SNR), which are the conjectural sources of VHE $`\gamma `$ rays; (ii) continuous studies of the physics of the relativistic jets in active galactic nuclei (AGN); (iii) further investigations of morphology and spectra of $`\gamma `$ rays from pulsar wind nebulae (PWN); (iv) the widening of the search for sources of pulsed $`\gamma `$-ray emission in VHE $`\gamma `$-ray band, etc. Such variety of physics enquiry can not be contented with only a single-type ground-based $`\gamma `$-ray instrument. Foremost the physics diversity of $`\gamma `$-ray emission mechanisms requires the essential observations appropriated in slightly different energy ranges. Thus for instance further observations of AGN and Pulsars necessitate the reduction of an instrumental energy threshold down to, at least, 10 GeV, whereas for detection of SNR assemblage a noticeable upgrade of sensitivity above 100 GeV is favored. Ultimately, the design of a major ground-based Čerenkov facility for future dedicated $`\gamma `$-ray observations has to conform to many requirements defined by peculiar energy spectral shapes, various angular extents, and strongly variable photon rates for the sources of an entirely different nature. ## 2 Basic Parameters of the Telescope Design The sensitivity of imaging atmospheric Čerenkov telescope (IACT) at an ascertained energy is mostly determined by the total amount of Čerenkov light photons, which the telescope is able to collect from the $`\gamma `$-ray showers of that specific primary energy. In general, the larger the number of photons in the Čerenkov light flash recorded from an individual atmospheric $`\gamma `$-ray shower, the higher the quality of the shower image. Three basic parameters account finally for a total number of Čerenkov light photons registered from the atmospheric shower by a telescope. First of all is the geometrical size of the reflector, $`A_o`$. It is apparent that the total number of recorded photons in a Čerenkov light flash scales linearly with respect to the geometrical area of the telescope’s reflector. A 10 m Whipple Čerenkov telescope has been operating without major complications for more than 30 years, which has indeed proven the robustness of a telescope of such size. The H.E.S.S. instrument comprises four 12 m telescopes. The challenge of achieving a pointing accuracy of about 20” and a point spread function of less than $`0.1^{}`$ have been achieved with H.E.S.S. (Bernloehr et al. 2002). MAGIC group recently started to perform $`\gamma `$-ray observations with a 17 m telescope built on La Palma, Canary Island. The development and construction of this telescope has allowed for a positive conclusion to the question of the technical feasibility of construction of a 30-35 m diameter Čerenkov telescope (Lorenz, Mirzoyan 2000). The optical performance of a telescope of such size has been also discussed by Hofmann (2001) and Akhperjanian, Sahakian (2003). Given all of those considerations one can indeed envisage that construction of a 30 m parabolic Čerenkov telescope is indeed practical (see Figure 1), considering both technical and financial aspects. Note that telescope reflectors of a diameter far beyond 30 m seem to be extremely costly. Moreover such a reflector will hardly meet the specifications for the off-axis point spread function, which is needed for a high-quality imaging of low energy $`\gamma `$-ray air showers. The Čerenkov light photons of an atmospheric shower which hit, along their transmission path, the telescope’s mirrored surface, can be captured by an advanced imaging camera placed in the focal plane of the reflector. Using modern PMs (photomultipliers) and fast electronics it now possible to convert registered photons into photoelectrons (ph.-e.), and finally digitize the output signal in a number of ADC (analog-to-digital converter) or FADC (flash analog-to-digital converter) counts. The overall efficiency of the photon-to-photoelectron conversion is another basic parameter of the telescope design. This efficiency, $`<ϵ>`$, can be calculated as $$<ϵ>=(\frac{1}{\lambda _1}\frac{1}{\lambda _2})^1_{\lambda _1}^{\lambda _2}r(\lambda )f(\lambda )q(\lambda )p(\lambda )e^{\sigma (\lambda ,d)}\lambda ^2𝑑\lambda $$ (1) where $`r(\lambda )`$ is the mirror reflectivity function, inferred from a set of measurements performed at different wavelengths; $`f(\lambda )`$ is the efficiency of light transmission of Winston cones, which are placed in front of the camera; $`q(\lambda )`$ is a quantum efficiency function of the PMs. The efficiency function, $`p(\lambda )`$, accounts for all remaining effects attenuating light, as for example, the shading the camera by the telescope masts etc. Usually it totals well below 10%. The wavelengthes, $`\lambda _1`$ and $`\lambda _2`$, are the boundaries of the wavelength range of detected Čerenkov photons, which lay within an optical band essentially between 0.3 and 0.6 $`\mu \mathrm{m}`$. For conventional telescopes the wavelength-averaged mirror reflectivity and the transmission of the Winston cones are about $`<r>0.8`$ and $`<f>0.7`$, respectively. Note that these efficiencies are close to the limit of the instrumental capacities and therefore they may not be substantially improved in the future. At the same time the quantum efficiency of the conventional PMs, which is of about $`q0.25`$ at maximum value, could be drastically improved for the modern advanced photo-detectors, e.g. PMs of APD S5345 type provide the quantum efficiency as high as $`q0.8`$ (see MAGIC Proposal, 1998). A substantial fraction of Čerenkov light photons emitted in atmospheric showers attenuates due to absorption and scattering of light onto the atoms, which Earth’s atmosphere is made up of. The corresponding reduction of the photon flux is given as $`\mathrm{\Phi }_a^{ph}e^{\sigma (\lambda ,d)}`$, where $`\sigma (\lambda ,d)`$ is a photon attenuation cross-section at the wavelength of $`\lambda `$, and for the effective distance of the photon emission point to the telescope, $`d`$. This effective distance is determined by the development height of an atmospheric shower, which depends on the primary energy of a shower, as well as by the altitude of the observational site, $`h_o`$. By setting the detector closer to the shower, which means up on the high altitudes, one can reduce the effective propagation length of the photons, $`d`$, in the atmosphere and correspondingly increase in the number of Čerenkov light photons arriving onto the telescope reflector. As Aharonian et al (2001) pointed out, the density of Čerenkov light photons at high altitude, $`h_o5`$ km above sea level, can increase, approximately, by a factor of 2 as compared with the corresponding density at conventional altitudes of $`h_o`$2 km above sea level. In fact, Čerenkov light photons are heavily absorbed whilst propagating within a few kilometers of the very dense atmospheric layer right above the observational level, which contains noticeable aerosol contamination. There are a number of high altitude sites in the world which are accessible for construction of the Čerenkov telescopes. However, the topology of the Čerenkov light images recorded at the high altitude site is such, that it makes implausible to achieve much of a gain in sensitivity from such an array located at high altitude (Konopelko 2004). Thus we have considered here a conventional observation level of $`h=1.8`$ km, which is close to the optimum one and eventually corresponds to the current H.E.S.S. site. Note that the H.E.S.S. collaboration has been funded to construct a 28 m telescope to be added to the current H.E.S.S. array, which will be place right in the center of the four telescope system. This telescope might be considered as a prototype for an array of five 30 m telescopes considered here. For the conventional PMs the overall photon-to-photoelectron conversion efficiency is about $`<ϵ>0.1`$, whereas using enhanced quantum efficiency photo-detectors one can achieve the conversion efficiency of $`<ϵ>0.40.5`$. However, such photo-detectors are still under development and are currently not commonly available on the market. Therefore, we considered here a telescope camera built using conventional PMs and electronics. Finally the basic parameters of the telescope’s design determine the effective area of the telescope reflector $`<A_o>=<ϵ>A_o`$, which is calculated with respect to a number of photoelectrons recorded by the camera. For the parameters chosen for a detector depicted here as STEREO ARRAY the corresponding effective collection area is about $`<A_o>70\mathrm{m}^2`$ (see Table 1). Ultimately, in the foreseeable future the large dish telescopes can be upgraded with cameras made of advanced photo-detectors, which will push the energy threshold as well as the sensitivity of this instrumentation to the absolute physical limit of the imaging atmospheric Čerenkov technique. ## 3 Energy Threshold The camera pixels of imaging atmospheric Čerenkov telescopes are able to function under a hard load of night sky background light, yielding a photoelectron rate of about 200 MHz per pixel. The camera trigger, which embraces the PM signals within a certain (trigger) zone of the camera, is designed so that it eliminates accidental triggers due to Poisson-like fluctuations of night sky background light over a large number of camera pixels. The optimum and robust trigger criterion allows for a reduction the rate of the accidental triggers down to a level which is at least one order of magnitude below the measured cosmic-ray rate. For a 30 m telescope such criterion requires, for instance, a simultaneous registration of about 3 pixels with a signal exceeding a $``$6 photoelectron level, which is finally determined by the telescope and camera design. In this case the minimal size (total number of photoelectrons in image) of the triggered events is around 20 photoelectrons. Events of such low size consist of only a few pixels and do not offer any reliable measurement of image shape and are in fact almost useless for stereoscopic shower reconstruction. Despite a few past trials, so far there is no well established analysis method that been developed specifically to extract such low size $`\gamma `$-ray events. In reality one has to apply a size cut at the level of 40 photoelectrons, which removes all images of low quality from the data sample on the costs of higher energy threshold. After applying selection by size the remaining events allow measurement of the orientation and the angular extensions along the major and minor axis of an ellipsoid-like image with sufficiently high accuracy. The minimal size of the recorded images is the final constraint on the actual energy threshold of the telescope. The density of the Čerenkov light photons emitted in the $`\gamma `$-ray-induced atmospheric shower, which was recalculated to a number of photoelectrons assuming specific telescope design, as a function of the distance of telescope to the shower axis in observational plane is shown in Figure 2. For a given reflector size and efficiency of photon-to-photoelectron conversion one can estimate corresponding minimal density at the telescope’s altitude, which corresponds to the minimal size cut. The lateral distribution of the photon density can be well described by a fit like $$\rho (R)=C(1+(R/R_{})^\beta )^1,$$ (2) where the fit parameters $`C,R_{},\beta `$ for a 100 GeV $`\gamma `$-ray shower are 14.0, 264.0, 3.41, respectively. One can see in Figure 2 that a broad plateau expanding up to the radius of R$``$120 m finally develops into an exponential fall-off<sup>1</sup><sup>1</sup>1This fit has a limited accuracy around R$``$120 m. It does not reproduce a very narrow bump, which occurs exactly at this radius due to high energy shower particles, but it is sufficiently good for the energy threshold estimate discussed here.. This lateral profile of Čerenkov light density resulted from a specific location of shower maximum in the atmosphere and a very narrow angle of Čerenkov light emission, of about 1. In a zero approximation a total amount of Čerenkov light photon emitted in a $`\gamma `$-ray shower is roughly proportional to the shower primary energy. Therefore, one can scale the density of Čerenkov light photons or size with respect to shower energy, $`E_{}`$, $$\rho (E_{},R)\rho (R)(E_{}/100\mathrm{GeV})^{1.2}.$$ (3) For instance for the density of Čerenkov light photons about $`\rho _{}\mathrm{\hspace{0.17em}0.6}\mathrm{ph}\mathrm{m}^2`$ the corresponding image size is of $`S=<A_o>\rho _{}40`$ ph.-e. One can see in Figure 2 that a 10 GeV $`\gamma `$-ray shower can be effectively detected by a 30 m telescope. At the same time a 1 TeV $`\gamma `$-ray shower at the impact distance of less than 250 m yields more than $`10^3`$ ph.-e., which is sufficient to produce a very high quality image. Note that the showers of a 10 TeV energy can be in principle detected at a 1 km impact distance, which results in a very large detection area being available for the stereo trigger, given a satisfactory large field of view of the imaging camera. Despite that various definitions of the energy threshold of an imaging atmospheric Čerenkov telescope have been discussed in the literature, at present it is widely accepted as a standard one the definition of the threshold as the energy, which corresponds to the position of the maximum in the differential detection rate of the $`\gamma `$-rays with the Crab-like energy spectrum in observations at zenith. The energy threshold of 500 GeV for the HEGRA system of five imaging atmopsheric Čerenkov telescopes estimated from the Monte Carlo simulations was fully confirmed by the direct measurements supported by a number of independent calibration tools (Konopelko et al. 1999). The same is true for the H.E.S.S. array of four telescopes in Namibia, which is right now in a state of routine data taking. The energy threshold of a 100 GeV predicted on the basis of Monte Carlo simulations (Aharonian et al. 1997) was verified recently using detailed systematic studies of the $`\gamma `$-ray fluxes from the Crab Nebula and other sources (Hinton, 2004). All that strengthen a confidence of the calculation of the energy threshold for a STEREO ARRAY given in this paper. ## 4 Focal Plane Detector Modern atmospheric Čerenkov telescopes focus the light from the atmospheric shower onto a fine-granularity imaging camera. An angular size of the pixel (PM) and the total field of view are the basic parameters of camera configuration. Optimum design of a camera must provide adequate registration of Čerenkov light images of $`\gamma `$-ray showers within a priori defined dynamic energy range. For the 30 m diameter telescopes the low energy bound is about 10 GeV. The shower maximum of such low energy $`\gamma `$-ray showers is high above the observational level \- $`H_{max}11`$ km. Therefore, the images of these showers have on average a very small angular size and in addition they are located very close to the center of the camera’s field of view. For an accurate measurement of angular size of these images one needs to reduce the angular pixel size down to about 0.07 in order to increase a number of tubes involved in image parametrization. At the same time small pixel size suppresses the contamination of the night sky background light accumulated within a signal readout time window (for detailed discussion see Konopelko, 2004). Note that further reduction of the pixel size is inexpedient, because such an extremely small pixel size will be beyond the reasonable limit given by optical performance of a $``$30 m parabolic dish (Hofmann 2001). In a toy model one can assume that most of Čerenkov light photons are coming directly from the shower maximum, $`H_{max}`$, which places at the atmospheric depth - $`p_{max}[\mathrm{gr}\mathrm{cm}^2]`$ \- which corresponds to the atmospheric depth of a maximum number of secondary electrons in a shower. If the telescope is tracking the source at the center of its field of view, then the position of the maximal image intensity (centroid) in the camera focal plane is determined then exclusively by the height of the shower maximum above the observation level and the geometrical distance of the shower axis to the telescope. In approximation A of the classical cascade theory, the shower maximum has taken place at the depth $$x=X_{}(ln(E_{}/E_c)1/2),\mathrm{gr}\mathrm{cm}^2$$ (4) where $`E_{}`$ is a primary energy of a $`\gamma `$-ray shower; $`E_c`$ is the so-called critical energy ($`E_c80`$ MeV), and $`X_{}`$ is a radiation length in atmosphere ($`X_{}=37.1\mathrm{gr}\mathrm{cm}^2`$). For a realistic atmospheric model the relation between the height, $`h`$ \[m\], and the atmospheric depth, $`x[\mathrm{gr}\mathrm{cm}^2]`$, can be given as $$h=(6740+2.5x)ln(1030/x).$$ (5) Finally the angular shift of the centroid can be calculated as $$\mathrm{\Theta }=\mathrm{arctan}[R(hh_{})^1],$$ (6) where $`h_{}`$ is a height of the observational level above sea level (hereafter $`h_{}1.8`$ km). For instance the shift of image centroid of a 10 GeV $`\gamma `$-ray shower at 200 m from the shower axis is about 1.25. Note that such toy model is accurate enough in most cases but it introduces a systematic error due to difference between the actual shower maximum position and the shower maximum as seen in the Čerenkov light, which is partially absorbed whilst propagating into the atmosphere. For a fixed primary energy Eqn. (3) can be inverted and the maximum radius, providing sufficient number of Čerenkov light photons for shower registration, can be calculated. Using Eqn. (6) one can estimate the corresponding angular shift of the image centroid in the camera focal plane. For example for a 1 TeV $`\gamma `$-ray shower the image displacement is about 1 at a 100 m impact distance<sup>2</sup><sup>2</sup>2Note that this case deals with a telescope pointing directly towards the source. However, the discussion given here also generally applies to other telescope pointings.. Bold curve in Figure 3 indicates the allowed range of the angular shifts for $`\gamma `$-ray showers of different primary energies. One can see in Figure 3 that for a 10 GeV energy $`\gamma `$-ray the image has a relatively small displacement from the center of field of view, which is below $`1.5^{}`$. The $`\gamma `$-ray showers of higher energies can trigger the detector at much large impacts if the angular size of the camera is sufficiently wide to accommodate these events. The point spread function of a 30 m parabolic reflector degrades relatively fast at $`1.5^{}`$ off the telescope optical axis. For example for a 17 m MAGIC telescope approximately 50% of all reflected light is concentrated within a circle of $`0.1^{}`$ for $`1^{}`$ off-axis observations (see MAGIC proposal). For a 30 m telescope assuming a realistic focal length (F/D=1.25) the point spread function will finally limit the angular range of efficient imaging by approximately $`1.52.0^{}`$ off the optical axis. For larger declinations of incoming photons the point spread function becomes substantially wider than the actual angular size of the image. The high energy $`\gamma `$-ray showers propagate deep into the atmosphere. Therefore their shower maximum occurs at relatively small heights above the observational level. If the telescope is pointing directly towards the source then the images of triggered high energy events will be concentrated in the camera focal plane at rather large angular distances from its center. An imaging camera with a limited angular size (field of view) of $`3.0^{}`$ diameter will be able to detect only Čerenkov light photons coming from the upper layers of the atmosphere, i.e. from the forefront of the shower development. This reduction of detection efficiency for very high energy $`\gamma `$-rays ($`E_{}1`$ TeV), due to a camera’s limited field of view, becomes even more severe for the telescopes deployed at high altitudes. In this case the shower maximum can be only observed at very large angles with respect to the telescope optical axis. There is always a trade-off between the angular size of the pixel and the total number of pixels in the camera, which are needed to cover a certain field of view. Cameras of a very small pixel size account for an enormous number of pixels/channels, which is, firstly, very expensive and secondly, very difficult in a long run exploitation. The pixel size of $`0.07^{}`$ for the field of view of 3 diameter is in fact a reasonable choice. It results in 1951 pixels, which is only by factor of two larger than that for the currently operating H.E.S.S. I cameras. An artist’s view of a possible camera design is shown in Figure 4. We decided to use the conventional ADCs for the signal digitization. At present the question of the advantageous performance of the FADCs for data acquisition and even for $`\gamma `$-ray imaging is still under general discussion and needs further verification. ## 5 System Layout The HEGRA (High Energy Gamma Ray Astronomy) experiment has proven for the first time numerous advantages of the stereoscopic observations of $`\gamma `$ rays above 500 GeV from the ground. H.E.S.S. (High Energy Stereoscopic System) was a natural extension of the HEGRA approach into lower energy range starting from 100 GeV. Utilizing a stereoscopic approach H.E.S.S. achieved the sensitivity at the level of $`1`$% of the Crab Nebula (standard candle) flux, which is so far unique in the field of very high energy $`\gamma `$-ray astronomy. Further reduction in energy thresholds down to $``$10 GeV also envisages making use of the stereoscopic approach. Large fluctuations in low energy showers noticeably degrade the quality of recorded Čerenkov light images. However, this can be substantially remedied using stereoscopic observations. To perform stereoscopic observations at low energies one needs apparently to have at least two 30 m imaging Čerenkov telescopes operating synchronically. Such a system of two telescopes might be considered as a basic stereoscopic system for observations of low energy $`\gamma `$ rays. Here we present the results of simulations for a 30 m single stand-alone telescope, as well as for a pair of such telescopes, and finally for a Stereo Array of five 30 m telescopes, which might be considered as a prototype for a future major facility in the field of very high energy $`\gamma `$-ray astronomy. The layout of the five telescopes is similar to that of the HEGRA layout but with a spacial separation between the telescopes of 100 m. A 100 m separation corresponds to a geometrical size of a Čerenkov light pool on the ground, which is in fact of about 100 m in radius. Despite the fact that a search for an optimum separation between two 30 m telescopes is an interesting issue, which needs perhaps special dedicated investigations, and might slightly influence the final sensitivity value, it is out of the frame of present paper, but it will be discussed in a separate forthcoming paper. ## 6 Simulations Atmospheric showers induced by $`\gamma `$-rays, electrons, and protons have been simulated using a numerical code described by Konopelko, Plyasheshnikov (2000). The primary energy of simulated showers was sampled as uniformly distributed within each of 24 bins, which were chosen within the energy range starting from 1 GeV and extending up to 10 TeV. Five 30 m Čerenkov telescopes were included in the simulational setup. The maximum impact distance of the shower axis to the center of the array was 1000 m. Basic parameters of the simulational setup are summarized in Table 2. The detailed procedure simulating the camera response was applied for the generated events. This procedure accounts for all possible losses of the Cherenkov light from the air shower on the way from the mirror reflecting to the camera response. The list of the effects which are important in this respect contains: (i) mirror reflecting sampled by the light raytracing technique or using the measured functions of the light spot distortion in the camera focal plane; (ii) PMT’s funnel acceptance; (iii) photon-to-photoelectron conversion inside the PMT’s taking into account the measured single photoelectron spectrum. Further details of the simulational procedure can be found in Konopelko et al. (1999). ## 7 Energy Spectra of Primary Particles In general a Čerenkov telescope counts events, which are atmospheric showers initiated by various particles of primary cosmic rays. Their major components are electrons, protons, and nuclei. The estimate for detector sensitivity is usually computed with respect to the well-established DC $`\gamma `$-ray flux from the Crab Nebula. The spectra of $`\gamma `$-rays and different background components measured in relevant energy range, which have been used in the sensitivity estimate given here, are summarized below. ### 7.1 Gamma-rays The DC flux of VHE $`\gamma `$-rays from the Crab Nebula is generally accepted as a standard flux unit in $`\gamma `$-ray astronomy. The spectrum of VHE $`\gamma `$-rays from the Crab Nebula was measured with a number of ground-based Čerenkov light detectors in energy ranges, starting from 50 GeV and extending up to 50 TeV (for review see e.g. Aharonian et al. 2000). The energy spectra and absolute fluxes measured by different instruments over different energy intervals are generally compatible. Therefore it is reasonable to use well-established spectrum of VHE $`\gamma `$-rays from the Crab Nebula for an estimate of the sensitivity of future Čerenkov light instruments and finally give a sensitivity merit expressed in Crab flux units. The Crab Nebula spectrum measured with the HEGRA system of 5 imaging atmospheric Čerenkov telescopes can be approximated by a power-law with a logarithmic steepening (Aharonian et al. 2000): $`dF_\gamma /dE=(2.67\pm 0.5)\times 10^{11}(\mathrm{E}/1\mathrm{TeV})^{2.47\pm 0.15(0.11\pm 0.10)\mathrm{log}(\mathrm{E})},`$ $`\mathrm{cm}^2\mathrm{s}^1\mathrm{TeV}^1`$ (7) The best fit indicates a slight flattening of the spectrum at the low energies, as is predicted by theories which assume an Inverse Compton (IC) origin of VHE $`\gamma `$ rays from the Crab Nebula (e.g. see de Jager et al. (1996); Atoyan & Aharonian (1999)). An extension of the HEGRA Crab Nebula spectrum to low energies, achieved with HEGRA by using a specific observational technique (Lucarelli et al. (2003)), is fully consistent with the recent measurements carried out with the Solar type detectors. Note that recent measurements of the Crab Nebula spectra in the energy range above 10 TeV (Aharonian et al. 2004) with HEGRA, and the preliminary data obtained with H.E.S.S. (Masterson et al. 2004) are in agreement. In fact the best fit of the HEGRA data, given by Eqn.(7), describes rather well all various spectral and flux measurements obtained so far. This fit was used here in computing the $`\gamma `$-ray detection rates. Note that the HEGRA and EGRET spectra of the DC $`\gamma `$-ray flux from the Crab Nebula are consistent within statistical and systematic errors given for both experiments. ### 7.2 Electrons For the ground-based atmospheric Čerenkov telescopes, which are approaching the energy range of $``$10 GeV, the electron component of primary cosmic rays becomes a dominant background and in fact is an issue of major concern. Around 10 GeV the intensity of the electron component is about $``$1% of the corresponding proton intensity. The Čerenkov light images generated by electron-induced atmospheric showers are almost indistinguishable in shape from the images of the $`\gamma `$-ray showers. Moreover the angular imaging resolution of Čerenkov telescopes substantially worsens at such low energies, and indeed does not allow for very strong suppression of an isotopic flux of cosmic electrons. Finally, the performance of imaging atmospheric Čerenkov telescopes in the energy range around 10 GeV relies heavily upon the residual trigger rate of electron-induced showers obtained after application of the analysis cuts. A number of various measurements of the energy spectrum of cosmic electrons have been performed in past (see Du Vernois et al. 2001). Despite that the existing data exhibit sizeable statistical and systematic uncertainties above 10 GeV the electron spectrum is a power-law with the spectral index of about 3.0. Over a broad energy range from 1 GeV up to 100 GeV the electron spectrum<sup>3</sup><sup>3</sup>3Hereafter we consider only the combined spectrum of $`e^+`$ and $`e^{}`$. can be fitted as $`F_e/dE=1.2\times 10^3\mathrm{E}^1(1+(\mathrm{E}/5\mathrm{GeV})^{2.3})^1,\mathrm{cm}^2\mathrm{sr}^1\mathrm{s}^1\mathrm{GeV}^1`$ (8) Such a steep energy spectrum makes it almost impossible to detect low fluxes of cosmic-ray electrons at high energies with currently operating IACTs, whereas absolute electron flux at energies of only a few GeV is relatively high. At low energies the flux of electrons becomes comparable with the flux of cosmic-ray protons after applying analysis cuts and it should be taken into account in total background estimate. Here we computed the electron rates using the spectral fit given by Eqn.(8). ### 7.3 Cosmic-Ray Protons & Nuclei The fluxes and spectra of primary cosmic-ray protons and helium nuclei can be thoroughly measured in the energy range of 1 to 200 GeV with the balloon born experiments. These spectra have been recently upgraded using the BESS spectrometer with the accuracies of $`5`$% for protons and $`10`$% for helium (Sanuki et al. 2000)<sup>4</sup><sup>4</sup>4In this paper a most complete compilation of modern data on differential proton and helium spectra is also given.. In the energy range of 1 to 100 GeV the 1998 BESS data on proton spectrum can be reproduced using the following fit $`dF_{\mathrm{CR}}/dE=0.3907\mathrm{E}^{0.3016}(1+(\mathrm{E}/1.776\mathrm{GeV}))^{3.141},`$ $`\mathrm{cm}^2\mathrm{s}^1\mathrm{str}^1\mathrm{GeV}^1`$ (9) Above 30 GeV the proton spectrum is a straight power law $$dF_{\mathrm{CR}}/dE=9.6\times 10^9(\mathrm{E}/10^3\mathrm{GeV})^{2.7},\mathrm{cm}^2\mathrm{s}^1\mathrm{str}^1\mathrm{GeV}^1.$$ (10) Note that the slope of a power law spectrum is consistent with the spectra given by Wiebel et al. (1994) and Simpson et al. (1983), which are $`dF_{\mathrm{CR}}/dEE^{2.75}`$ and $`dF_{\mathrm{CR}}/dEE^{2.66}`$, respectively, whereas the absolute fluxes differ by 20-50%. Since at present BESS measurements offer the best quality data, the proton spectrum described by Eqns.(9,10) was used here in calculations of the cosmic-ray detection rates. It is known that Helium nuclei in primary cosmic rays contribute up to $``$30% to the total raw event rate of atmospheric Čerenkov telescopes (see e.g. Konopelko et al. 1999). The contribution of all other cosmic-ray nuclei is very low and almost negligible. That was proven here by specific test simulations. In a good approximation one can scale the event rate calculated for the proton cosmic ray flux by a factor of 1.5 in order to get a total cosmic ray detection rate. ## 8 Performance of STEREO ARRAY An estimate of the detector performance is inferred from the detection rates of $`\gamma `$-ray and background events, calculated after application of the analysis cuts. For a given flux of cosmic-ray particles of a specific type the energy dependent function of correspondent detector response, i.e. collection area, basically determines relevant detection rate. Collection areas for $`\gamma `$-ray- and cosmic-ray- induced showers can be derived from the detailed Monte Carlo simulations tracing both the shower development in the atmosphere and the detector performance. Below we summarize the basic results on collection areas of $`\gamma `$ rays and background events of a different kind. ### 8.1 Collection Areas In the simulations the position of a shower axis in the observational plane was sampled as uniformly randomized over an area limited by 1 km radius from the center of the array. By counting the number of triggered showers of a specific primary energy one can calculate a so-called trigger probability, which then can be easily converted into corresponding collection area. Events were simulated within the energy range from 1 GeV up to 10 TeV. The energy dependence of the collection area basically reflects the profile of the Čerenkov light pool on the ground. For $`\gamma `$-ray showers it has a high density plateau expanding up to 120 m distance from the shower axis, which then develops into an exponential fall-off. One can see in Figure 5 that a steep rise of collection area at low energies (E$``$50 GeV) converges into a rather slow logarithmic growth at high energies. For $`\gamma `$-ray showers the characteristic energy at which the energy profile of collection area breaks roughly corresponds to the energy threshold of the instrument. For proton showers the Čerenkov light pool does not have such a prominent “shoulder” at 120 m but it yields rather smoothly decreasing photon density with enlargement of the distance to the shower core. The collection areas of $`\gamma `$-ray atmospheric showers can be consummately reproduced by a four parameter fit: $$S_\gamma =a_1E^{a_2}(1+(E/a_3)^{a_4})^1,\mathrm{cm}^2,$$ (11) The parameters of the fit for different detector configurations are given in Table LABEL:parm1. The proton-induced showers are isotropically distributed over the incidence angle with respect to the telescope optical axis. Thus in this case the collection area has to be computed by averaging over a solid angle, limited by an opening angle, which exceeds by roughly $`1^{}`$ the opening angle of the camera’s field of view. Interestingly, the second parameter of the fit, $`a_3`$ (Eqn.(11)), roughly corresponds to the energy threshold of the instrument with respect to the corresponding type of atmospheric showers, which is about 8 GeV for $`\gamma `$-ray showers and 40 GeV for proton-induced showers. The trigger efficiency for atmospheric showers of some specific energy is finally determined by a total amount of Čerenkov light emitted in the shower of such an energy (see discussion above). For low energy proton-induced showers ($`E<`$100 GeV) less than 30% of their primary energy can be transferred into electromagnetic component, which is finally responsible for the Čerenkov light emission. That explains why the energy thresholds for the $`\gamma `$-ray- and proton-induced atmospheric showers differ greatly. The development of electron-induced showers in the atmosphere is almost identical to the $`\gamma `$-ray-induced showers. A slight difference in a height of shower maximum for the two cases does not noticeably effect the lateral and angular parameters of Čerenkov light pool at the observational level. In observations of a point-like $`\gamma `$-ray source the isotropically distributed background events of cosmic-ray protons and electrons are registered within a solid angle, which is limited by actual angular resolution of the instrument. The angular resolution strongly depends on primary energy of $`\gamma `$-ray shower (see below). For STEREO ARRAY even at very low energies, i.e. around the instrument’s energy threshold of $``$10 GeV, the angular resolution is still rather good, $`\mathrm{\Theta }0.3^{}`$. It is substantially less than the angular size of the camera, which is of $`3^{}`$ diameter. Deviation of the incident angles of electron-induced showers within $`0.3^{}`$ from the telescope optical axis does not affect the physics of a shower development in the atmosphere, namely the lateral and longitudinal profiles of secondary shower particles etc. Therefore, it is possible to use the collection areas calculated for the $`\gamma `$-ray-induced showers in calculating the collection areas of electron-induced showers. Additional test simulations of some electron-induced atmospheric showers have proven such approximation to be valid. In Figure 5 the collection areas for a system of five telescopes for different trigger multiplicities are also shown. It is important to note that the $`\gamma `$-ray events of energy below 10 GeV trigger at most only two of the telescopes. The 3-fold coincidences are generated by $`\gamma `$ rays of noticeably higher energy. Finally, the collection area of 5-fold coincidence events dominates at the energy above 30 GeV. In fact, the low energy $`\gamma `$-ray showers can trigger a telescope at quite small impact distances to the shower axis, which explains the energy dependence of the collection areas for different trigger multiplicities. ### 8.2 Detection rates Differential detection rates for different types of atmospheric showers registered with STEREO ARRAY can be calculated as $$dR_{(\gamma ,e,CR)}/dE=(dF_{(\gamma ,e,CR)}/dE)S_{(\gamma ,CR,e)}\mathrm{Hz}\mathrm{GeV}^1$$ (12) The position of a peak in the differential detection rate of $`\gamma `$-ray atmospheric showers, which is calculated assuming the Crab Nebula energy spectrum, defines the effective energy threshold of the instrument (see Figure 6). Note that it is, in fact, in agreement with the value of the $`a_3`$-parameter of a fit of the collection area (see Table LABEL:parm1). One can see in Figure 6 that the electron-induced showers dominate at low energies, whereas their contribution is negligible at high energies due to the very steep energy spectrum of cosmic ray electrons. The differential detection rate of the proton-induced atmospheric showers peaks at substantially higher energy, within 30-90 GeV. The proton-induced atmospheric shower of that energy yields approximately the same amount of Čerenkov light as the $`\gamma `$-ray showers of about 10 GeV. The integral raw background rate for a system of two 30 m telescopes is expected to be about 1 kHz. At the same time the raw detection rates for a single 30 m telescope, as well as for a system of five such telescopes, are expected to be of 1.7 and 3.2 kHz, respectively. With up-to-date electronics one can achieve a sufficiently low dead-time-of-event processing, and it is generally possible to maintain such high event rate with an advanced imaging camera. However the discussion of possible solutions to the problems of trigger electronics and data acquisition is beyond the framework of present studies. ### 8.3 Analysis The raw observational data taken with the ground-based atmospheric Čerenkov telescopes are entirely background dominated. The detection rate of the background cosmic rays and electrons usually exceeds by three orders of magnitude the expected rate from a Crab like $`\gamma `$-ray source. A detailed comparative analysis of space, angular, and temporal characteristics of 10 GeV and 100 GeV $`\gamma `$-ray showers is given in Konopelko (1997). The images of the sub-100 GeV $`\gamma `$-ray showers are indeed very irregular in shape and are strongly affected by high fluctuations in air shower development. However they still carry information about shower orientation as well as its lateral, and longitudinal spread. In order to extract the tiny fraction of the $`\gamma `$-ray events from the enormous amount of background contaminations one has to apply specific analysis cuts. The set of the analysis cuts (selecting criteria) usually consists of the selections on the image orientation and shape. Two telescopes make it possible to perform full-scale stereo reconstruction of the shower direction in space and on the ground. #### 8.3.1 Orientation: The deflection of the reconstructed arrival direction of the $`\gamma `$-ray showers from the nominal source position is used for their orientational selection. The better the angular resolution of the instruments, the lower the content of residual background events. The angular resolution (angular radius limiting the area around the source position accounting for 63% of all simulated showers) as a function of primary energy of the $`\gamma `$-ray-induced atmospheric shower is shown in Figure 7. At low energies the Čerenkov light images contain less light and are strongly affected by shower fluctuations. Therefore the results on reconstruction of the image orientation as well as the shower arrival direction substantially worsen. For the angular size of the camera of $`\mathrm{\Theta }_o1.5^{}`$ in radius the cosmic ray rejection is about $`\eta 0.04`$. One can see in Figure 7 that the angular resolution achieved with three images is 30% better than in the case of only two images. It allows the reduction of the cosmic ray background contamination by a factor of 1.7. However, as already mentioned above, 3-fold triggers substantially decrease the rate of $`\gamma `$ rays of energy close to the threshold. That is why to search for $`\gamma `$-rays at the energy of 10 GeV and below the 2-fold coincidence mode seems to be more advantageous. #### 8.3.2 Image shape: Exploitation of the HEGRA system of 5 imaging atmospheric Čerenkov telescopes with H.E.S.S. allowed us to conclude that the most effective parameter of cosmic ray discrimination by the image shape is a parameter of mean scaled Width (m.s.w.). At least with two telescopes operating in a stereoscopic mode one can reconstruct the impact distance of the shower axis from each of the telescopes, and scale the actual transverse angular size of the image with the total amount of light in the image ($`Size`$-parameter) at that impact distance. The distributions of the parameter of mean scaled Width for the simulated $`\gamma `$-ray and proton showers are shown in Figure 8. Despite a significant overlap between two distributions one can reduce the content of the proton showers by factor of $`10`$ (i.e. the ”after cuts” acceptance of cosmic ray showers is $`\kappa _{CR}=0.08`$) applying mean scaled Width cut of 0.91. This cut keeps 35% of all registered $`\gamma `$-ray showers. The corresponding quality-factor is Q-factor=$`\kappa _\gamma (\kappa _{CR})^{1/2}1.2`$. High fluctuations in the image $`Size`$ of low energy events cause such rather modest rejection power. Simulations have shown that an application of any other additional parameter of image shape does not improve the resultant quality factor. In observations of a point like $`\gamma `$-ray source the recorded images of the $`\gamma `$-ray-induced air showers are located in the camera focal plane mostly within the angular annulus limited by inner and outer radii of 0.3 and 1.2 from the camera center, respectively<sup>5</sup><sup>5</sup>5Note that due to the high rate of the low energy events they are basically dominant in the entire sample of registered $`\gamma `$-rays.. Thus before applying the orientation and shape cuts one can noticeably enhance the content of the $`\gamma `$-ray showers by using the $`Distance`$ cut (angular distance of the image centroid to the camera center). Such selection criterion allows for the reduction of the rate of the background events by a factor of 2. Since the shape of the images generated by the electron- and $`\gamma `$-ray-induced atmospheric showers are almost identical, one can in both cases use the acceptances of the analysis cuts derived for the $`\gamma `$-ray-induced showers. The analysis of data taken with a single stand-alone telescope at shower energies around 10 GeV is not straight-forward. Note that there were a number of trials using advanced, sometimes very sophisticated, analysis methods to improve the rejection of the background and consequently to enhance the $`\gamma `$-ray abundance in data taken with a single telescope. However none of them demonstrated drastically improved performance as compared with the standard multi-parametric approach (”super-cuts”). Recent studies on the multi-variate analysis in application to the data simulated for a 30 m telescope have shown that a realistic quality-factor for such low energy events is of about $`Q3`$, which corresponds to the $`\gamma `$-ray and cosmic ray acceptance of $`\kappa _\gamma 0.3`$ and $`\kappa _{CR}0.01`$, respectively (Konopelko et al. 2005). This was achieved using Bayesian decision rules and a non-parametric estimate of the multi-parameter density. The analysis cuts for a single telescope as well as for a system of telescopes were selected here primarily on the basis of equal $`\gamma `$-ray acceptance across the entire dynamic energy range. Therefore these cuts do not affect noticeably the position of the peak in the differential detection rate and the estimated value of the energy threshold. ### 8.4 Sensitivity The sensitivity of the Čerenkov telescope in observations of the point-like $`\gamma `$-ray sources can be defined in different ways. One approach, which is widely used, is to calculate the minimum flux of the $`\gamma `$ rays, which can be detectable after 50 hrs of observations at the significance level above 5$`\sigma `$ ($`\sigma `$ is one standard deviation of a number of background events). The total number of $`\gamma `$-ray, proton, and electron showers after applying analysis cuts, which remains in the data sample of observational time $`t`$, can be calculated as $$N_{(\gamma ,p,e)}=t_{E_0}^E_{\mathrm{}}(dJ_{(\gamma ,p,e)}/dE)\stackrel{~}{S}_{(\gamma ,p,e)}(E)\kappa _{(\gamma ,p,e)}^d\kappa _{(\gamma ,p,e)}^s𝑑E$$ (13) where $`E_0`$ is the selected<sup>6</sup><sup>6</sup>6Here the selection of the energy threshold is defined by the physics of the source, which finally defines the favorable energy range. energy threshold energy. $`E_{\mathrm{}}`$ is the upper energy boundary, which is chosen well above the actual threshold of the detector ($`E_{\mathrm{}}`$ is usually the maximal energy of the simulated showers, and $`E_{\mathrm{}}10^3E_{th}`$) allowing accurate calculation of the detection rates. The detection areas of proton and electron showers are collection areas averaged over the solid angle $`\mathrm{\Omega }_o2\pi \theta _{max},\theta _{max}=4^{}`$. Thus the collection areas used in Eqn. (13) are $`\stackrel{~}{S}_\gamma =S_\gamma `$, and $`\stackrel{~}{S}_{(p,e)}=<S_{(p,e)}>\mathrm{\Omega }_o`$. $`\kappa _{(\gamma ,p,e)}^{(d,s)}`$ are the acceptances of the $`\gamma `$-ray-, proton-, and electron-induced showers after applying directional and shape cuts, respectively. The integral detection rates for the $`\gamma `$-ray, electron, and proton showers are summarized in Table 4. The corresponding signal-to-noise ratio after 50 hrs exposure can be calculated as $$(S/N)=N_\gamma (N_p+N_e)^{1/2}$$ (14) In order to derive the $`\gamma `$-ray flux, which yields the desirable significance of 5$`\sigma `$ one can scale Crab flux as $$F^{min}(>E_0)=5(S/N)^1F^{Crab}(>E_0)$$ (15) For the real detection of the source one has to control in addition a number of detected $`\gamma `$-ray showers, which should not be less than 10, which sets the second condition of the minimum detectable flux $$F^{min}(>E_0)10(N_\gamma )^1F^{Crab}(>E_0)$$ (16) Note that for the detector considered here the first condition is always stronger, which means that the total number of detected $`\gamma `$-rays is in fact high, and the second condition is almost always holds ($`N_\gamma >>10`$). The final results on the sensitivity calculations for the system of two 30 m telescopes are summarized in Table 4. It is worth noting that at low energies the electron background is dominant over the cosmic ray background. The electron component of the primary cosmic rays at energies about 5 GeV becomes, in fact, a major background. The images of electron and $`\gamma `$-ray showers are almost indistinguishable in shape, which make the electron background an absolute limiting factor for the detector sensitivity. Improved angular resolutions might help to reject contamination of the isotropic electron flux in observations of the point-like $`\gamma `$-ray source. The sensitivity of the system of two 30 m telescopes can be compared to the corresponding sensitivities of the single stand-alone telescope of the same size as well as with the array of five such telescope operating in the stereoscopic mode - STEREO ARRAY. The signal and background rates, as well as resulting sensitivities are given in Table 5<sup>7</sup><sup>7</sup>7Note that for any of given configurations the peak in the differential detection rate lays well above the chosen energy threshold – $`E_o`$=5 GeV.. One can see in Table 5 that single stand-alone telescopes provide rather high $`\gamma `$-ray rates. However, due to limited rejection power the final sensitivity is by more than a factor of 2 lower than for a system of two telescopes, and by a factor of 5 as compared with the system of five telescopes. At the same time the increase in sensitivity with two to five telescopes is relatively modest. It is provided by the increase in collection area of the $`\gamma `$-ray showers, as well as a slightly better angular resolution for higher telescope multiplicity events. Given the results obtained here one can consider an alternative approach, which is a construction of a few systems of two 30 m telescopes spatially separated in the observational plane. In such a case the sensitivity of entire telescope array will be proportional to $`n^{1/2}`$, where $`n`$ is a number of independent two telescopes subsystems. ### 8.5 Array Flexibility The flexibility in operation of the array of the atmospheric imaging Čerenkov telescopes is in fact an important issue. For instance one can use two subsystems of the array for simultaneous observations of two different sources. Apparently it increases the observational time by factor of two, but with a somewhat reduced sensitivity. On other hand in some cases all telescopes of the array can be used independently in a stand-alone mode for the monitoring of a number of potential $`\gamma `$-ray sources (like AGNs) to perform quick-look discovery observations for the sample of promising candidates. ## 9 Conclusion Based on the detailed Monte Carlo simulations we have studied the performance of the system of two, as well as five, 30 m imaging atmospheric Čerenkov telescopes. An array of five 30 m telescopes is very close to the optimum detector for the $`\gamma `$-ray observations above 10 GeV. Such an array of detectors allows for the reduction of the energy threshold down to a few GeV, and would be the most sensitive instrument in the field of ground-based high energy $`\gamma `$-ray astronomy in this energy range. The STEREO ARRAY might be considered as a prototype for a major future detector for ground-based high energy $`\gamma `$-ray astronomy. Its construction and operation will enable scientists to perform high quality $`\gamma `$-ray observations from the ground, which will provide an overwhelming insight into the understanding of the mechanics of high energy $`\gamma `$-ray emission from the various physical environments in the Universe. The physics rationale of the STEREO ARRAY, provided with a sensitivity value as given here, is a subject for another forthcoming paper. ## Acknowledgements I would like to thank James Buckley, Werner Hofmann, and Trevor Weekes for the discussions on a subject of this paper. I would like also thank the referee, who remains anonymous, for comments and suggestions, which have improved the quality of the paper. ## References 1. Aharonian et al. The energy spectrum of TeV gamma rays from the Crab Nebula as measured by the HEGRA system of imaging air Čerenkov telescopes, ApJ, 539, 317, (2000) 2. Aharonian et al. The Crab Nebula and Pulsar between 500 GeV and 80 TeV: Observations with the HEGRA Stereoscopic Air Cerenkov Telescopes, ApJ, Volume 614, Issue 2, pp. 897 (2004) 3. Aharonian, F., Hofmann, W., Konopelko, A., Völk, H.J. The potential of the ground based arrays of imaging atmospheric Cherenkov telescopes, Astroparticle Physics, Volume 6, Issue 3-4, 343; 369; (1997) 4. Aharonian, F., Konopelko, A., Völk, H.J., Quintana, H.5@5 - a 5 GeV energy threshold array of imaging atmospheric Cherenkov telescopes at 5 km altitude, Astroparticle Physics, Vol. 15, N4, 335 (2001) 5. Akhperjanian, A., Sahakian, V. Performance of a 20 m diameter vCerenkov imaging telescope, Astroparticle Physics, Volume 21, Issue 2, p. 149 (2004) 6. Atoyan, A., Aharonian, A. On the mechanisms of gamma radiation in the Crab Nebula, MNRAS, 278, 525 (1996) 7. Bernloehr et al. The optical system of the H.E.S.S. imaging atmospheric Cherenkov telescopes, Astroparticle Physics, Vol. 20, Issue 2, 111 (2003) 8. de Jager et al. Gamma-Ray Observations of the Crab Nebula: A Study of the Synchro-Compton Spectrum, ApJ, 457, 253 (1996) 9. DuVernois et al. Cosmic-ray electrons and positrons from 1 to 100 GeV: measurements with HEAT and their interpretation, ApJ, 559, 296 (2001) 10. Hinton, J. The status of the HESS project, New Astronomy Reviews, Vol. 48, Issue 5-6, 331 (2004) 11. Hofmann, W. How to focus a Cherenkov telescope, J. Phys. G: Nucl. Part. Phys. 27, No 4, 933 (2001) 12. Konopelko, A., et al. Performance of the stereoscopic system of the HEGRA imaging air Cerenkov telescopes: Monte Carlo simulations and observations, Astroparticle Physics, Volume 10, Issue 4, p. 275 (1999) 13. Konopelko, A. Space-angular and temporal parameters of Cherenkov light emission in air showers of energy from 1 TeV down to 10 GeV, Proc. of the Kruger Park Workshop on TeV Gamma-Ray Astrophysics ”Towards a Major Atmospherics Cherenkov Detector - V, (ed. O.C. de Jager), Kruger Park, South Africa, August 8-11, 208(1997) 14. Konopelko, A. Altitude effect in Cerenkov light flashes of low energy gamma-ray-induced atmospheric showers, Journal of Physics G: Nuclear and Particle Physics, Volume 30, Issue 12, pp. 1835 (2004). 15. Konopelko, A., Chlingarian, A., Reimer, A., Study on background rejection and energy resolution with a 30 m stand-alone imaging atmospheric Cherenkov telescope using non-parameteric multi-variate methods in a sub-100 GeV energy range, in preparation (2005) 16. Lorenz, E., Mirzoyan, R. What have we learned from the MAGIC telescope developments for a future large IACT ?, Proc. Intern. Symp. “High Energy Gamma-ray Astronomy”, Eds. F.Aharonian and H.J. Völk, AIP Conf. Proceed. 558, 586 (2000) 17. Lucarelli, F., Konopelko, A., Aharonian, F., Hofmann, W., Kohnle, A., Lampeitl, H., Fonseca, V. Observations of the Crab Nebula with the HEGRA system of IACTs in convergent mode using a topological trigger, Astroparticle Physics, Vol. 19, Issue 3, 339 (2003) 18. MAGIC Proposal: “The MAGIC Telescope”: design study for the construction of a 17 m Čerenkov telescope for gamma-ray astronomy above 10 GeV MPI-PhE/98-5 (1998) 19. Masterson, C. Observation of galactic TeV gamma ray sources with H.E.S.S., Proc. 28th Int. Cosmic Ray Conf., Tsukuba, Univ. Academy Press, Tokyo, p. 2323 (2003) 20. Sanuki et al. Precise measurement of cosmic-ray proton and Helium spectra with the BESS spectrometer, ApJ, 545, 1135 (2000) 21. Simpson, J.A. Ann. Rev. Nucl. and Part. Sci., 33, 323 (1983) 22. Wiebel, B. Chemical composition in high energy cosmic rays, Fachbereich Physik Bergische Iniversität Gesamthochschule, Wuppertal, WUB 94-08 (1994)
warning/0506/gr-qc0506018.html
ar5iv
text
# Classification of Cohomogeneity One Strings ## I Introduction Recently, extended objects, such as strings and membranes, gather much attention in the contexts of unified theories and cosmologies. A trajectory of the extended object, world hypersurface, is a submanifold embedded in a target space. The geometry induced on the world hypersurface has indeed much variety which depends on the possible solutions to the equations of motion of the object, but extended objects which have geometrical symmetries are especially interesting for their simplicity. A matter of main interest in the brane universe scenario is the structure of symmetric extended objects, i.e., spatially homogeneous brane universe models embedded in a symmetric bulk spacebrane . Another example is classical stationary solutions for Nambu-Goto string in stationary spacetimes Frolov:1988zn ; Carter:1989bs ; Frolov:1996xw ; Vilenkin-Shellard . Equations of motion of an extended object governed by the Nambu-Goto action or its generalization has the form of partial differential equations with non-linear constraint equations in general. However, the equations reduce to ordinary differential equations when the object has enough symmetries. In the investigations noted above, rich properties of the extended objects are clarified by solving the ordinary differential equations explicitly. Suppose that an $`m`$-dimensional hypersurface is embedded in an $`n(>m)`$-dimensional target space which has isometries, the embedded hypersurface can possess a part of the symmetries of the target space. If a subgroup of the isometries of target space acts on the hypersurface, the restriction on the hypersurface is isometries of it. In addition, the extrinsic geometry of the hypersurface is also symmetric with the isometries. The hypersurface is said to be a cohomogeneity one hypersurface if a subgroup of isometries of a target space acts on the hypersurface, and the orbits of the action on it span an $`(m1)`$-dimensional space. As is discussed later in the Nambu-Goto string case, the equations of motion for a cohomogeneity one object can be reduced to ordinary differential equations. Among a variety of extended object’s dimension, we concentrate ourself especially on cohomogeneity one strings, as the simplest extended objects, in this letter. The one-parameter group of isometries acting on the world surface is generated by a Killing vector field on a target space, that is, the string’s world surface is tangent to the Killing vector field. If a target space has only one Killing vector, there is one class of cohomogeneity one strings associated with the Killing vector. On the other hand, in a space with more than two independent Killing vectors, there would be infinite classes of cohomogeneity one strings since infinite numbers of linear combinations of the Killing vectors are possible. Let us consider the four-dimensional Minkowski spacetime, which has ten linearly independent Killing vectors, as a target space. Suppose two Killing vectors which denote translation symmetries along different directions. Though these Killing vectors are linearly independent, cohomogeneity one strings associated with these Killing vectors are members of the same class because the world surfaces of these strings are transformed each other by a suitable isometry of Minkowski spacetime which rotates one direction of translation symmetry into the other. The issue treated in this letter is how many classes of cohomogeneity one strings are there in a symmetric spacetime. For clarifying this problem, we classify the Killing vectors in Minkowski spacetime, as a typical example of symmetric spacetime, introducing an equivalence relation by using isometries of the spacetime. ## II Equations of Motion of Cohomogeneity One String Let us consider a two-dimensional world surface $`\mathrm{\Sigma }`$ is embedded in an $`n`$-dimensional target spacetime $``$ with a metric $`g_{ab}`$ which possesses Killing vector fields. If the world surface $`\mathrm{\Sigma }`$ is tangent to one of the Killing vector fields of $``$, say $`\xi ^a`$, we call the world surface $`\mathrm{\Sigma }`$ a cohomogeneity one string associated with the Killing vector $`\xi ^a`$. The stationary string is one of the example associated with a Killing vector which is timelike on the surfaceFrolov:1996xw . When we choose a Killing vector field in $``$, there would be many cohomogeneity one world surfaces associated with the Killing vector. If the equations of motion governing the dynamics of the string is given, the possible cohomogeneity one string surfaces are selected out as solutions to the equations. Hereafter, we consider the Nambu-Goto string, as the simplest example, whose action is given by $`S={\displaystyle _\mathrm{\Sigma }}\sqrt{\gamma },`$ (1) where $`\gamma `$ is the determinant of induced metric on the world surface. The orbit space of a Killing vector field $`\xi ^a`$ of $``$, say $`𝒩`$, is an $`(n1)`$-dimensional space on which the metric $`h_{ab}=g_{ab}\xi _a\xi _b/f`$ (2) is introduced naturally, where $`f:=\xi ^a\xi _a`$. The metric $`h_{ab}`$ has Euclidean signature in the region where $`f<0`$ and Lorentzian signature in $`f>0`$. Suppose that a curve $`𝒞`$ is given in $`𝒩`$, orbits of the Killing vectors $`\xi ^a`$ starting from $`𝒞`$ span a two-dimensional surface. Since $`h_{ab}`$ measures the length along the direction perpendicular to $`\xi ^a`$, area of the surface element is simply given by $`dA=\sqrt{|f|}\sqrt{|h_{ab}dx^adx^b|},`$ (3) where $`dx^a`$ is an infinitesimal displacement along $`𝒞`$. Therefore, the action (1) reduces to $`S={\displaystyle _𝒞}\sqrt{\stackrel{~}{h}_{ab}dx^adx^b},`$ (4) where $`\stackrel{~}{h}_{ab}:=fh_{ab}.`$ (5) The action (4) gives the length of $`𝒞`$ with respect to the metric $`\stackrel{~}{h}_{ab}`$ on $`𝒩`$. Therefore, the problem for finding solutions of cohomogeneity one string associated with $`\xi ^a`$ reduces to the problem for solving $`(n1)`$-dimensional geodesic equations with respect to the metric $`\stackrel{~}{h}_{ab}`$. When a world surface is associated with a spacelike Killing vector, i.e., $`f>0`$ on the surface, $`𝒞`$ should be a timelike geodesic with respect to the metric $`\stackrel{~}{h}_{ab}`$ so that $`\mathrm{\Sigma }`$ is a timelike world surface. ## III Classification in Minkowski Spacetime First, let us consider a pair of two-dimensional world surfaces, $`\mathrm{\Sigma }_A`$ and $`\mathrm{\Sigma }_B`$, embedded in a symmetric target space $``$ which admits isometries. The world surfaces $`\mathrm{\Sigma }_A`$ and $`\mathrm{\Sigma }_B`$ are geometrically equivalent if there is an isometry $`\varphi `$ of $``$ which maps $`\mathrm{\Sigma }_A`$ onto $`\mathrm{\Sigma }_B`$: $`\varphi :\mathrm{\Sigma }_A\mathrm{\Sigma }_B.`$ (6) Next, let $`\mathrm{\Sigma }_A`$ be a cohomogeneity one string associated with a Killing vector field $`𝝃_A`$ and let $`\mathrm{\Sigma }_B`$ be a geometrically equivalent to $`\mathrm{\Sigma }_A`$. The isometry $`\varphi `$ which maps $`\mathrm{\Sigma }_A`$ to $`\mathrm{\Sigma }_B`$ pushes forward $`𝝃_A`$ to a vector field, say $`𝝃_B`$: $`\varphi ^{}:𝝃_A𝝃_B.`$ (7) It is clear that $`𝝃_B`$ is also a Killing vector field on $``$, and $`\mathrm{\Sigma }_B`$ is a cohomogeneity one string associated with it. If $``$ has a large number of symmetries, it is possible that $`𝝃_A`$ and $`𝝃_B`$ are linearly independent vector fields which are connected by the isometry $`\varphi `$. Therefore, it suggests the idea that we classify Killing vector fields by the equivalence relation with respect to the isometries. We introduce an equivalence relation, $``$, as follows: Killing vectors $`𝝃_A`$ and $`𝝃_B`$ are equivalent if and only if there is an isometry $`\varphi `$ which pushes forward $`𝝃_A`$ to $`\lambda 𝝃_B(\lambda :\text{constant})`$: $`\varphi ^{}:𝝃_A\lambda 𝝃_B𝝃_A𝝃_B.`$ (8) The scalar multiplication comes from the fact that the Killing vector field is irrelevant to the constant scaling. Here, we classify the Killing vectors in Minkowski spacetime based on the equivalence relation noted above. In Minkowski spacetime with the metric $`ds^2=dt^2+dx^2+dy^2+dz^2,`$ (9) there are ten independent Killing vectors which generate Poincaré group: $`𝑷_a(a=t,x,y,z),`$ Translations along $`a`$-direction; $`𝑲_i(i=x,y,z),`$ Lorentz boosts along $`i`$-direction; $`𝑳_i(i=x,y,z),`$ Space rotations around $`i`$-axis. An arbitrary Killing vector $`𝝃`$ is expressed as a linear combination of them: $`𝝃=\alpha _a𝑷_a+\beta _i𝑲_i+\gamma _i𝑳_i,`$ (10) where $`\alpha _a`$, $`\beta _i`$ and $`\gamma _i`$ are constant coefficients. We will list up inequivalent representatives under the equivalence relation (8) using Poincaré group of isometries. The task we should do is logically straightforward but tedious. Then, we would like to show some typical procedure and final result. As an example of the procedure of classification, we consider a Killing vector in the form $`𝝃=a𝑲_y+b𝑳_z,`$ (11) where $`a`$ and $`b`$ are arbitrary constants. We use the Lorentz boost along $`x`$-direction denoted by $`\varphi =\text{Exp}[\phi 𝑲_x]`$, where $`\phi `$ is a parameter. The Killing vector $`𝝃`$ is pushed forward by $`\varphi `$ as $`\varphi ^{}:𝝃(a\mathrm{cosh}\phi b\mathrm{sinh}\phi )𝑲_y+(a\mathrm{sinh}\phi +b\mathrm{cosh}\phi )𝑳_z.`$ (12) Choosing the parameter $`\phi `$ suitable for the values of $`a`$ and $`b`$, we see that $`𝝃`$ is equivalent to one of three Killing vectors: $`𝝃=`$ $`a𝑲_y+b𝑳_z\{\begin{array}{cc}𝑲_y\hfill & \text{for }|a|>|b|\hfill \\ 𝑳_z\hfill & \text{for }|a|<|b|\hfill \\ 𝑲_y+𝑳_z\hfill & \text{for }|a|=|b|\hfill \end{array},`$ (13) where we also use space or time reflection. In contrast, it is worth while to note that the Killing vector in the form $`𝝃=a𝑲_z+b𝑳_z,`$ (14) for example, is not equivalent to $`𝝃^{}=a^{}𝑲_z+b^{}𝑳_z`$ (15) except the case $`|a/b|=|a^{}/b^{}|`$. In the metric (9), the norm of $`𝝃`$ is given by $`a^2(z^2+t^2)+b^2(x^2+y^2),`$ (16) Then, equinorm surfaces of $`𝝃`$ are in the form of squashed hyperboloids, where $`a^2/b^2`$ describes the amount of squashing. Since the pushing forward $`\varphi ^{}`$ preserves the norm of vectors then the shape of the equinorm surfaces should be same for equivalent vector fields. If $`|a/b||a^{}/b^{}|`$ the amount of squashing of the hyperboloid is different. Thus, there is no isometry which maps $`𝝃`$ to $`𝝃^{}`$, i.e., $`𝝃`$ is inequivalent to $`𝝃^{}`$. Starting from a general linear combination (10), we can complete the classification. We find that equivalence classes are partitioned into seven families listed in Table I. Each partition consists of infinite classes parameterized by one parameter. We introduce a notation, $``$, such that $`𝑷_t𝑳_z`$, for example, means the set of all linear combinations $`a𝑷_t+b𝑳_z`$, where $`a`$ and $`b`$ are non-negative constants satisfying $`a^2+b^2=1`$. The constraints on $`a`$ and $`b`$ are for eliminating total scaling of Killing vectors and redundancy under space and time reflections. The number of partitions would crucially depends on the structure of the group of isometries. Since Poincaré group is the group of isometries preserving indefinite Minkowski metric then the first three families, I$``$ III, should be distinguished. If the classification is done in Euclidean space, these families would fall into one family since the rotations in $`tz`$ plane connect members in these families. Similarly, families of III, IV and V would fall into one family in the flat Euclidean space. In addition, it seems that the existence of translation group makes the issue complicated. It would be interesting to classify the Killing vectors in other maximally symmetric spacetime: de Sitter and anti-de Sitter spacetimesdesitter . The Killing vector fields in the families I, V, VI and VII can be timelike in some regions in Minkowski spacetime except edges of the families i.e., $`𝑳_z,𝑷_z`$, and $`𝑲_y+𝑳_z`$. Then, there are four types of stationary strings which correspond to four families of timelike Killing vectors. The rigidly rotating strings are one of them; they are associated with the Killing vectors in the family $`𝑷_t𝑳_z`$. All Killing vector fields except $`𝑷_t`$ and $`(𝑷_t+𝑷_z)`$ have spacelike region then there are seven types of cohomogeneity one strings with spacelike symmetries. From (2) and (5), we see that a one-parameter family of Killing vectors gives one-parameter family of metrics $`\stackrel{~}{𝒉}`$ on the orbit space $`𝒩`$. Then, we should solve seven types of geodesic equations for finding all of cohomogeneity one Nambu-Goto string solutions in Minkowski spacetime. Some of them are easily solved Frolov:1996xw Ogawa:etal . Generalizations of the present work to higher dimensional target spaces and higher dimensional cohomogeneity one objects are interesting issues. It is easy to get ordinary differential equations for higher dimensional cohomogeneity one objects, but classification of them would require some labors. ###### Acknowledgements. We are grateful to Dr. K. Nakao and Dr. Y. Yasui for helpful discussions. This work is supported by the Grant-in-Aid for Scientific Research No.14540275.
warning/0506/cond-mat0506142.html
ar5iv
text
# A large magnetic storage ring for Bose-Einstein condensates ## Abstract Cold atomic clouds and Bose-Einstein condensates have been stored in a $`10`$cm diameter vertically-oriented magnetic ring. An azimuthal magnetic field enables low-loss propagation of atomic clouds over a total distance of $`2`$m, with a heating rate of less than $`50`$nK/s. The vertical geometry was used to split an atomic cloud into two counter-rotating clouds which were recombined after one revolution. The system will be ideal for studying condensate collisions and ultimately Sagnac interferometry. The field of atom optics erlchaifan has seen a plethora of significant advances since the advent of laser cooling lascool and gaseous Bose-Einstein condensation (BEC) And1 . Analogs of mirrors, lenses, and beamsplitters for manipulating ultra-cold atoms now exist and are continually being improved. A relatively recent addition to the atom-optical toolbox is the cold atom storage ring, where atoms are guided around a closed path. This is an interesting configuration for performing, for instance, ultra-sensitive Sagnac Sag atom interferometry. An electrostatic storage ring was first reported for molecules Mei in 2001. However, for many interferometry experiments atoms are more suitable candidates, as they can be easily laser cooled and/or prepared in the same quantum state. The first atomic storage ring was formed with the magnetic quadrupole field created by two concentric current carrying loops sau . Our storage ring us utilises the more symmetric quadrupole field of a four-loop geometry (Fig. 1). A third group has recently developed a storage ‘stadium’ prent based on a magnetic waveguide. Our ring contains more than $`5\times 10^8`$ atoms, has a lifetime of $`50`$s and an area of $`72\mathrm{cm}^2,`$ each corresponding to more than an order of magnitude improvement with respect to Refs. sau ; prent . The Sagnac effect Sag is linearly proportional to the area of an interferometer, and our ring’s area compares favourably with that of a state-of-the-art thermal beam atom interferometer gyro $`(A=0.22\mathrm{cm}^2`$ thermgyro ). The atoms can complete multiple revolutions in the ring, further increasing our effective area and thus the rotation sensitivity. In addition, our storage ring has the unique feature that we are able to form a BEC in a section of the ring us , and observe its propagation around the ring. We begin by considering the storage ring theoretically, before discussing our experimental setup. Comparisons will then be drawn between our experiment and theory. All magnetic atom-optical elements make use of the Stern-Gerlach potential $`U=\mu _Bg_Fm_FB`$ experienced by an atom moving adiabatically in a magnetic field of magnitude $`B`$, where $`\mu _B`$ is the Bohr magneton, $`m_F`$ the atom’s hyperfine magnetic quantum number, and $`g_F`$ is the Landé g-factor. To make a storage ring one must use atoms in weak-field-seeking magnetic states $`(g_Fm_F>0),`$ which are attracted to minima of the magnetic field magnitude. Using the Biot-Savart law, one can express the cylindrically symmetric magnetic field from a single coil of radius $`R,`$ $`\text{b}_R(r,z),`$ in terms of elliptic integrals good . Our storage ring comprises four concentric circular coils (Fig. 1), with a toroidal quadrupole total magnetic field: $`\text{B}(r,z)`$ $`=`$ $`\text{b}_{R\delta _R}(r,z\delta _z)\text{b}_{R\delta _R}(r,z+\delta _z)`$ (1) $``$ $`\text{b}_{R+\delta _R}(r,z\delta _z)+\text{b}_{R+\delta _R}(r,z+\delta _z).`$ In our experiment $`R=5.0\mathrm{cm},`$ $`\delta _R=1.25\mathrm{cm},`$ and $`\delta _z=1.35\mathrm{cm},`$ leading to a ring of zero magnetic field at a radius $`R_0=4.8\mathrm{cm}`$ slightly smaller than the mean coil radius $`R.`$ The axial wire (Fig. 1) adds a $`1/r`$ azimuthal magnetic field to Eq. 1, which has little effect on the ring radius, but yields a storage ring with non-zero magnetic field. The theory of the storage ring for cold atoms is relatively simple, as the spatio-temporal evolution of each atom can be accurately described using the equations of motion arising from the Stern-Gerlach potential. After choosing suitable Gaussian initial position and velocity distributions it is possible to perform a Monte Carlo simulation to build up a spatio-temporal atomic density map. By comparison with a 3D model, we have found that a 1D model is sufficient, i.e. the rigid pendulum equation: $$\theta ^{\prime \prime }(t)=(g/R_0)\mathrm{sin}\theta ,$$ (2) where $`\theta `$ and $`\omega =\frac{d\theta }{dt}`$ are the angular position and velocity of an atom, $`g`$ is the acceleration due to gravity and $`R_0`$ is the radius of the storage ring. Note that Eq. 2 can be integrated analytically (using $`\frac{d^2\theta }{dt^2}=\omega \frac{d\omega }{d\theta })`$ given the initial angle $`(\theta _0)`$ and angular velocity $`(\omega _0)`$ of an atom, to determine the time-averaged relative probability of finding an atom at a given angle: $$P(\theta )1/\omega =1/\sqrt{\omega _{0}^{}{}_{}{}^{2}+2g(\mathrm{cos}\theta _0\mathrm{cos}\theta )/R_0}.$$ (3) In our vertically-oriented storage ring there is therefore a high time-averaged probability of finding atoms in the region where they travel the slowest – the top of the ring. The integral of Eq. 3, leads to an expression for $`t(\theta )`$ in terms of an elliptic integral, which can be inverted to find $`\theta (t)`$. There are two kinds of trajectories in the ring: if $`\mathrm{cos}(\theta _{max})=|\frac{R_0}{2g}\omega _{0}^{}{}_{}{}^{2}+\mathrm{cos}\theta _0|>1`$ an atom will always rotate in the same direction, but if $`\mathrm{cos}(\theta _{max})1`$ (cf. Fig. 2(b)) the atom will reverse its direction around the ring at the turning points $`\theta =\pm \theta _{max}`$. The theoretical time-dependent angular distribution of atoms can be seen in Fig. 2. An atomic cloud centered in a parabolic potential can expand or shrink, but will maintain the same cloud shape. A circular atomic trajectory has a potential which is only parabolic to second order, and it is the higher order effects which cause an atomic cloud released from the top of the ring ($`\theta `$=0) to break into two halves (Fig. 2(a)), leaving a near-zero probability at the top of the ring after around $`300`$ms. These two halves return to the top of the ring after a further $`300`$ms and have a non-zero probability of remaining at the top of the ring for all subsequent times. The time for the two halves of the atomic cloud to ‘recombine’ has only a weak dependence on the initial atomic temperature. Gravity is the dominant effect for cold atoms, leading to a ‘hot’ velocity of $`1.4\mathrm{m}\mathrm{s}^1`$ at the bottom of the ring – a good location for future studies of high-energy ($`20`$mK) collisions between BECs WilWal . Although the spread of initial thermal velocities quickly ensures that atoms reach the top of the ring at different times, in Fig. 2(a) we still see slowly decaying ‘echos’ of the originally localised spatial atomic cloud in the form of time-varying bimodal perturbations of the $`\theta `$ distribution. The angular width of these echos is approximately proportional to the initial cloud width, so the echos and their bimodal nature become more pronounced at higher temperatures. In many of our ring experiments the atoms/BECs were not launched from $`\overline{\theta _0}=0`$ (Fig. 2(a)). Although the total number of atoms observed is similar to the $`\overline{\theta _0}0`$ case, the size of the returning $`\overline{\theta _0}0`$ cloud is much smaller and clearer due to focusing at the turning points of its motion (Fig. 2(b)). We now turn to the experimental setup, for which many of the details are in Refs. us ; AARIIS . Our double magneto-optical trap (MOT) dmot collects $`10^9`$ <sup>87</sup>Rb atoms in both our high and low pressure MOT chambers. The storage ring can be loaded directly from the low pressure MOT or Ioffe-Pritchard (IP IP ) trap without the need for any additional magnetic guiding or transfer, i.e. the MOT and IP trap are at the top of the ring. Releasing atoms at the top of a vertical ring ensures insensitivity to magnetic ring corrugations and a fast rotation frequency. The advantage of our hybrid magnetic trap (Fig. 1) is that, using the same coils, it can be (i) a MOT, (ii) an IP magnetic trap (with trap frequencies $`\nu _r=230\mathrm{Hz},`$ $`\nu _z=10\mathrm{Hz})`$ or iii) a toroidal magnetic storage ring with radial gradient $`230`$G/cm. In modes (i) and (ii) the magnetic field has an adjustable aspect ratio in the azimuthal direction. The four 2 turn$`\times 500`$A circular coils in Fig. 1 create a toroidal magnetic quadrupole field, confining the atoms to a ring. The four 3 turn$`\times 500`$A square ‘pinch’ coils are wired in pairs, and confine the atoms to a localised section of the ring for Ioffe-Pritchard trapping of the atoms. An azimuthal field can be added to the storage ring via an axial wire. The large coil sizes ensure good optical access, and low magnetic field noise. The absorption imaging (gravity) direction is the $`x`$ $`(y)`$ axis. After loading the low pressure MOT, the atomic cloud is optically pumped, loaded into the IP trap, magnetically compressed, and then evaporatively cooled to an adjustable temperature. The atoms can then be smoothly loaded into the storage ring in $`20`$ms. We are able to form Bose-Einstein condensates containing $`N_0=2\times 10^5`$ $`|F=2,m_F=2`$ atoms at the top of the storage ring us at a typical final RF evaporation frequency of $`750\mathrm{kHz}`$. In order to compare our experimental ring data (Fig. 3) with our theoretical model (Fig. 2), we have found it convenient to model the total number of atoms in the ‘viewing window’ of our absorption imaging system as a function of time. Our CCD camera has an area of $`4.8\times 6.4\text{mm}^2,`$ with a magnification of $`1.20(1).`$ We have released cold atomic clouds with a variety of final evaporation temperatures into a storage ring with no azimuthal field. In accordance with theory, the atoms disappear after $`300400`$ms, reappear shortly afterwards, and are present at all subsequent times. However, there is a marked variation in the storage ring dynamics with atomic cloud release temperature. This can be explained in terms of non-adiabatic Majorana spin-flip transitions TOP . The storage ring has a ring of zero magnetic field, and colder atoms pass closer to the magnetic field zero, and are selectively removed from the storage ring. If we apply a constant magnetic field across the entire storage ring, perpendicular to the ring axis, it is possible to have only two places in the ring with zero magnetic field, however there will be a strong angular variation in the trapping potential. A novel feature of our storage ring is that the ‘hole’ in our quartz vacuum chamber (Fig. 1) allows us to use a wire along the axis of the storage ring. This means that we can generate an azimuthal magnetic bias field $`B_\theta =010`$G around the ring which transforms the radial magnetic potential from a cone to a hyperbola, removing the ring of zero magnetic field and drastically reducing atomic loss. This difference can be seen in the experiments of Fig. 3(a) and (b), where $`2\mu `$K atomic clouds are released into a ring without and with an azimuthal field, respectively. A quantitative contrast of the number of visible atoms in each case is shown in Fig. 4, as well as a comparison to the Monte Carlo theory of Fig. 2. The difference the azimuthal field makes is even greater with condensates: without it BECs vanish before completing one revolution, but with an azimuthal field multiple revolutions of a condensate are possible with low loss (Fig. 3(c)). Note that when we performed experiments with an azimuthal field, the atoms/BECs were not launched from $`\overline{\theta _0}=0.`$ If atoms are launched from $`\overline{\theta _0}=0`$ (Fig. 2(a)), then although the total number of atoms observed (Fig. 4) is similar to the $`\overline{\theta _0}0`$ case, the size of the returning $`\overline{\theta _0}0`$ cloud is much smaller and clearer due to focusing in the ring (Figs. 2(b), 3(b,c)). We do not expect to see phase-fluctuations phasefluc in our condensate before or after propagation in the ring. These effects have been studied in highly elongated condensates in which the BEC coherence length is less that the length of the condensate. We form the condensate in only a moderately elongated trap, and the azimuthal expansion process is rapid enough that the (density-dependent) phase fluctuations do not have time to develop. Sagnac interferometry will be performed by locating the condensate at the exact top of the ring, and incoherently splitting the sample by simply releasing it (Fig. 5). We are currently looking for interference fringes after a single revolution. Coherent splitting of the BEC will be achieved using Bragg scattering bragg to send BEC wavepackets in both directions around the ring. Note that ideally the phase sensitivity of a BEC interferometer scales like $`\delta \varphi _{BEC}N^1`$ where $`N`$ is the number of atoms (cf. $`\delta \varphi _TN^{1/2}`$ for thermal atoms), however the increased sensitivity is only possible if the BECs are prepared in number states yama . In conclusion, we have demonstrated a storage ring for cold <sup>87</sup>Rb atomic clouds and BECs with an area of $`7200\mathrm{mm}^2`$. An azimuthal bias field around the ring enables low-loss BEC propagation with heating of less than $`50`$nK/s. Our goal is a highly sensitive Sagnac atom interferometer, in which we are aided by our unprecedented ring area. Rotation sensitivity for a single revolution of $`\mathrm{\Delta }\mathrm{\Omega }=\mathrm{}/(8m\pi R_{0}^{}{}_{}{}^{2}\sqrt{N})=\mathrm{3\hspace{0.17em}10}^{11}\mathrm{rad}/\mathrm{s}`$ is feasible. If one cannot create an azimuthal magnetic field, it is still possible to prevent spin-flip losses by using an adjustable-radius time-orbiting ring trap (TORT) TORT . We recently learnt that a team in Berkeley has created the first TORT for BECs stamper , with ring area $`5\mathrm{mm}^2`$. Large BECs were also formed in an optically-plugged ring trap with ring area $`1\mathrm{mm}^2`$ raman . We are grateful for helpful discussions with K. Burnett and E. Hinds. This work was supported by the UK EPSRC and the University of Strathclyde.
warning/0506/hep-ph0506173.html
ar5iv
text
# The equivalence of fluctuation scale dependence and autocorrelations ## 1 Introduction Fluctuations in nuclear collisions measured at a single bin size or scale could arise from many possible configurations of a multiparticle momentum distribution and are therefore difficult to interpret. However, the scale dependence of fluctuations over a significant scale interval does contain detailed information about multiparticle correlations which can be extracted with the proper techniques . Information about the absolute location of event-wise structure is lost, but those aspects depending only on position difference are retained in the form of autocorrelation distributions . In this paper we consider fluctuations on binned momentum space, Pearson’s correlation coefficient and autocorrelation density ratios for multiplicity $`n`$ and transverse momentum $`p_t`$. Autocorrelations can be inferred directly from pair ratios or from a fluctuation/autocorrelation integral equation which we derive. Inverting the integral equation we obtain autocorrelations which can be compared with more conventional conditional distributions. ## 2 Fluctuations and correlations on binned spaces Correlation analysis of nuclear collisions reveals information arising from event-wise changes in multiparticle momentum distributions. The data system is an ensemble of event-wise particle distributions on momentum space $`(p_t,\eta ,\varphi )`$ or $`(y_t,\eta ,\varphi )`$, where $`p_t`$ is transverse momentum, $`m_t`$ is transverse mass, $`\eta `$ is pseudorapidity, $`\varphi `$ is azimuth and $`y_t\mathrm{ln}\{(m_t+p_t)/m_0\}`$ is transverse rapidity with pion mass assigned to $`m_0`$. The momentum space is bounded by a detector acceptance, and the space within the acceptance is binned according to one or more bin sizes. Particle number $`n`$ is distributed on the full momentum space. It is useful to think of transverse momentum $`p_t`$ or rapidity $`y_t`$ as a continuous measure distributed on angle subspace $`(\eta ,\varphi )`$ and sampled by individual particles. Number $`n`$ and transverse momentum $`p_t`$ or rapidity $`y_t`$ correlations can be considered both separately and in conjunction. An ensemble of event-wise histograms on momentum subspace $`x`$ is represented schematically in Fig. 1 (first panel), with bins $`a,b`$ singled out. Events can be compared with the ensemble-mean distribution to determine relative information, measured by fluctuations of bin contents about their means. The ensemble mean of event-wise pair distributions on space $`(x_1,x_2)`$ in Fig. 1 (second panel) can be compared to a reference distribution consisting of cartesian products of single-particle mean distributions. The difference reveals correlations in the two-particle distribution corresponding to fluctuations in the single-particle distribution . That relation is the basis for the integral equation connecting fluctuations to correlations described below. Fig. 1 (third panel) sketches possible frequency distributions on number combinations $`(n_a,n_b)`$ from bin pair $`(a,b)`$ in the first panel. The ellipses represent half-maximum contour lines for gaussian-random fluctuations. The three cases correspond to correlation (solid curve), anticorrelation (dash-dot curve) and no correlation (dashed curve) between bins, the last being expected from a mixed-pair or central limit reference. The 2D frequency distribution is characterized by two marginal variances and a covariance. The marginal variances for bins $`a`$ and $`b`$ are given by $`\sigma _{a,b}^2=\overline{(n\overline{n})_{a,b}^2}=\overline{n_{a,b}^2}\overline{n}_{a,b}^2`$. The covariance between those bins is given by $`\sigma _{ab}^2=\overline{(n\overline{n})_a(n\overline{n})_b}=\overline{n_an_b}\overline{n}_a\overline{n}_b=\sigma _\mathrm{\Sigma }^2\sigma _\mathrm{\Delta }^2`$, where $`\sigma _\mathrm{\Sigma }^2`$ and $`\sigma _\mathrm{\Delta }^2`$ are variances along sum and difference diagonals in the third panel. Pearson’s correlation coefficient is a measure of relative covariance . For bin pair $`(a,b)`$ it is defined by $`r_{ab}\sigma _{ab}^2/\sqrt{\sigma _a^2\sigma _b^2}=\{\sigma _\mathrm{\Sigma }^2\sigma _\mathrm{\Delta }^2\}/\{\sigma _\mathrm{\Sigma }^2+\sigma _\mathrm{\Delta }^2\}[1,1]`$. The numerator is the ($`a,b`$) covariance and the denominator is the geometric mean of the marginal $`a`$ and $`b`$ variances. That coefficient is our basic correlation measure. Quantities $`r_{ab}`$, determined for all histogram bin pairs in the second panel, completely represent fluctuations on space $`x`$. Variances and covariances depend on the bin size or scale on $`x`$. The scale dependence of fluctuations is in turn directly related to two-particle correlations, as described in this paper. ## 3 Object and reference distributions We distinguish between an object distribution, part of whose correlation content we wish to measure, and a reference distribution which contains by construction information in the object distribution we wish to ignore. We determine object and reference distributions for number $`n`$ correlations and $`p_t`$ correlations (the latter require a slightly different treatment, as described below). One single-particle reference for fluctuation measurements is the ensemble-mean histogram on $`x`$. Variances and covariances for single bins and bin pairs measure the average information in the event-wise object distribution relative to the ensemble mean. For two-particle distributions on $`(x_1,x_2)`$, object pair density $`\rho _{obj}(x_1,x_2)`$ is constructed from sibling pairs taken from same events, and reference pair density $`\rho _{ref}(x_1,x_2)`$ is constructed from mixed pairs taken from different but similar events. The reference could also be a Cartesian product of single-particle means (therefore factorizable by construction). Object and reference distributions are combined in several ways: 1) density of correlated pairs: $`\mathrm{\Delta }\rho (x_1,x_2)\rho _{obj}(x_1,x_2)\rho _{ref}(x_1,x_2)`$; 2) density of correlated pairs per particle pair: $`\mathrm{\Delta }\rho (x_1,x_2)/\rho _{ref}(x_1,x_2)\rho _{obj}(x_1,x_2)/\rho _{ref}(x_1,x_2)1`$; 3) density of correlated pairs per particle: $`\mathrm{\Delta }\rho (x_1,x_2)/\sqrt{\rho _{ref}(x_1,x_2)}\{\rho _{obj}(x_1,x_2)\rho _{ref}(x_1,x_2)\}/\sqrt{\rho _{ref}(x_1,x_2)}`$. 1) and 2) are conventional correlation measures. 3) is unconventional, but closely related to Pearson’s correlation coefficient described in the previous section: a relative covariance measure which does not depend on the absolute number of particles per se. Pearson’s coefficient is ideally suited for testing linear superposition in the context of heavy ion collisions. ## 4 Autocorrelations The autocorrelation concept was first introduced to time-series analysis in the form $`A(\tau )1/T_{T/2}^{T/2}f(t)f(t+\tau )𝑑t`$, where $`\tau `$ is the lag . The concept is most useful when function $`f(t)`$ is stationary: its correlation structure does not depend on absolute location on time. The information in $`f(t)`$ is then fully represented by the autocorrelation distribution on $`\tau `$. The autocorrelation concept can be generalized to spatial correlations. If event-wise structure is randomly positioned on space $`x`$ then the corresponding ensemble-average two-point distribution on $`(x_1,x_2)`$ is stationary (not depending on absolute position on $`x`$). Distributions in Fig. 2 (left two panels) of measure $`\mathrm{\Delta }\rho (x_1,x_2)/\rho _{ref}(x_1,x_2)`$ on $`\eta `$ and $`\varphi `$ are typical of Au-Au collisions at RHIC . Two-particle distributions can also be defined on sum and difference variables: $`\rho (x_1,x_2)\rho (x_\mathrm{\Sigma },x_\mathrm{\Delta })`$, with $`x_1,x_2x_\mathrm{\Sigma }x_1+x_2,x_\mathrm{\Delta }x_1x_2`$. The data distributions in Fig. 2 exhibit stationarity—they do not depend on sum variable $`x_\mathrm{\Sigma }`$—in which case we have $`\rho (x_\mathrm{\Sigma },x_\mathrm{\Delta })\rho (x_\mathrm{\Delta })`$ to good approximation. We then average the two-particle density over $`x_\mathrm{\Sigma }`$ to obtain the autocorrelation density on $`x_\mathrm{\Delta }`$ (still a 2D density, not a projection). The autocorrelation on difference variables can be constructed in two ways as shown in Fig. 2 (last two panels): 1) bin space $`x`$ with microbins of size $`ϵ_x`$ and average the bin contents of space $`(x_1,x_2)`$ along diagonals, as discussed further below, 2) bin difference variable $`x_\mathrm{\Delta }`$ in space $`(x_1,x_2)`$ directly and form the corresponding pair histogram . In both cases, care must be taken to insure that true averages are obtained and not projections by integration. We define autocorrelation densities $`\rho (x_\mathrm{\Delta })`$, autocorrelation histograms $`A_k(ϵ_x)ϵ_x^2\rho (2kϵ_x)`$ and joint autocorrelations on two or more difference variables. If the primary distribution on $`x`$ is truly stationary, the autocorrelation is a lossless compression of the two-particle momentum distribution to a lower-dimensional space. ## 5 Autocorrelation density ratio for number correlations We now define a universal correlation measure for nuclear collisions. Returning to Pearson’s correlation coefficient we make the following approximation $`r_{ab}{\displaystyle \frac{\sigma _{ab}^2}{\sqrt{\sigma _a^2\sigma _b^2}}}={\displaystyle \frac{\overline{(n\overline{n})_a(n\overline{n})_b}}{\sqrt{\overline{(n\overline{n})_a^2}\overline{(n\overline{n})_b^2}}}}{\displaystyle \frac{\overline{(n\overline{n})_a(n\overline{n})_b}}{\sqrt{\overline{n}_a\overline{n}_b}}}={\displaystyle \frac{\overline{n_an_b}\overline{n}_a\overline{n}_b}{\sqrt{\overline{n}_a\overline{n}_b}}}.`$ (1) We replace the marginal number variances in the denominator of $`r_{ab}`$ by their Poisson values. The result is the histogram equivalent of density ratio $`\mathrm{\Delta }\rho (x_1,x_2)/\sqrt{\rho _{ref}(x_1,x_2)}`$ previously defined. This relative covariance can be interpreted as the number of (anti)correlated pairs per particle (explicit in the last expression). The third combination of object and reference distributions in Sec. 3 and the modified Pearson’s normalized covariance in Eq. (1) are equivalent. This density ratio is the basic correlation measure for any bin pair $`(a,b)`$ on space $`x`$. We then define an autocorrelation in terms of density ratios for sets of bins $`(a,b)`$ on $`(x_1,x_2)`$: we combine density ratios as averages along diagonals $`k`$, assuming the basic distribution on $`(x_1,x_2)`$ is approximately stationary. The average on index $`a`$ along the $`k^{th}`$ diagonal in Fig. 2 (third panel) is $`{\displaystyle \frac{\mathrm{\Delta }A_k(n)}{\sqrt{A_{k,ref}(n)}}}\left\{{\displaystyle \frac{\overline{(n\overline{n})_a(n\overline{n})_{a+k}}}{\sqrt{\overline{n}_a\overline{n}_{a+k}}}}\right\}_{\overline{a}}ϵ_{x_\mathrm{\Delta }}{\displaystyle \frac{\mathrm{\Delta }\rho (n;kϵ_{x_\mathrm{\Delta }})}{\sqrt{\rho _{ref}(n;kϵ_{x_\mathrm{\Delta }})}}}.`$ (2) This is the autocorrelation definition for analysis of nuclear collisions on angle space $`(\eta ,\varphi )`$. This density ratio is an intensive correlation measure which precisely measures relative correlations even for excursions of object and reference densities over orders of magnitude. ## 6 Extension to transverse momentum $`p_t`$ correlations We now extend the definition of the density ratio and its autocorrelation to distributions of transverse momentum $`p_t`$ on $`(\eta ,\varphi )`$. Measurement of $`p_t`$ fluctuations is described in . We treat $`p_t`$ as a continuous measure distributed on space $`x`$, with scalar sums of particle $`p_t`$ in histogram bins. We could write the Pearson’s coefficient for transverse momentum by analogy with number correlations in terms of difference $`(p_t\overline{p}_t)`$. However, the corresponding per-particle $`p_t`$ variance can be expressed as the sum of three terms: $`\overline{(p_t\overline{p}_t)^2}/\overline{n}=\overline{(p_tn\widehat{p}_t)^2}/\overline{n}+2\widehat{p}_t\overline{(p_tn\widehat{p}_t)(n\overline{n})}/\overline{n}+\widehat{p}_t^2\overline{(n\overline{n})^2}/\overline{n}`$, a ‘$`p_t`$’ variance, a $`p_t`$-$`n`$ covariance and a number variance. The three terms have forms similar to Pearson’s normalized covariance, but with $`a=b`$ defining corresponding normalized variances (also called ‘scaled variances’). Each term is important in the overall problem of particle and $`p_t`$ production. We therefore want to describe the structure of event-wise $`p_t`$ distributions in terms of $`(p_tn\widehat{p}_t)`$, independent of but coordinated with the structure of number distributions described in terms of $`(n\overline{n})`$. Pearson’s correlation coefficient for transverse momentum correlations thus takes the form $`r_{ab}{\displaystyle \frac{\overline{(n\overline{n})_a(n\overline{n})_b}}{\sqrt{\overline{n}_a\overline{n}_b}}}{\displaystyle \frac{\overline{(p_tn\widehat{p}_t)_a(p_tn\widehat{p}_t)_b}}{\sigma _{\widehat{p}_t}^2\sqrt{\overline{n}_a\overline{n}_b}}},`$ (3) i.e., a relative $`p_t`$ covariance as opposed to a relative number covariance. The geometric mean of marginal variances in the denominator is in this case replaced by $`\sigma _{\widehat{p}_t}^2\sqrt{\overline{n}_a\overline{n}_b}`$, the mean of central-limit expectations for the $`p_t`$ variances. The factor $`\sigma _{\widehat{p}_t}^2`$ is however omitted in what follows to be consistent with the first term of the per-particle $`p_t`$ variance expansion above Eq. (3). Factors $`\sigma _{\widehat{p}_t}^2`$ or $`\widehat{p}_t^2`$ may be introduced as necessary in a subsequent interpretation stage. The corresponding $`p_t`$ density-ratio autocorrelation is $`{\displaystyle \frac{\mathrm{\Delta }A_{kl}(p_t:n)}{\sqrt{A_{kl,ref}(n)}}}\left\{{\displaystyle \frac{\overline{(p_tn\widehat{p}_t)_{ab}(p_tn\widehat{p}_t)_{a+k,b+l}}}{\sqrt{\overline{n}_{ab}\overline{n}_{a+k,b+l}}}}\right\}_{\overline{ab}}ϵ_{\eta _\mathrm{\Delta }}ϵ_{\varphi _\mathrm{\Delta }}{\displaystyle \frac{\mathrm{\Delta }\rho (p_t:n;kϵ_{\eta _\mathrm{\Delta }},lϵ_{\varphi _\mathrm{\Delta }})}{\sqrt{\rho _{ref}(n;kϵ_{\eta _\mathrm{\Delta }},lϵ_{\varphi _\mathrm{\Delta }})}}}.`$ (4) That expression is formulated explicitly in terms of 2D correlations on $`(\eta ,\varphi )`$. An analogous 2D expression can be derived for number correlations from Eq. (2). Fig. 3 shows joint autocorrelations of per-particle density ratios on angle difference variables for number correlations (first two panels) and $`p_t`$ correlations (second two panels) and for charge-independent (like-sign $`+`$ unlike-sign) pair combinations (left panel of each pair) and charge-dependent (like-sign $``$ unlike-sign) pair combinations (right panel of each pair) . Contributions from self pairs were excluded. These joint autocorrelations, determined directly by pair counting, provide full access to all angular correlations on $`(\eta ,\varphi )`$ but are computationally expensive for larger event multiplicities. We can also obtain these autocorrelations by inverting fluctuation scale dependence, but with much less computation effort. ## 7 Relating fluctuations and correlations Fluctuation scale dependence results from event-wise correlation structure in single-particle distributions which can be extracted by solving an integral equation. We first define differential scale-dependent fluctuation measures based on variance differences for number $`n`$ and transverse momentum $`p_t`$ fluctuations, then derive the integral equation connecting fluctuations and correlations. Number variance difference $`\mathrm{\Delta }\sigma _{n/}^2(\delta x)\overline{(n(\delta x)\overline{n}(\delta x))^2}/\overline{n}(\delta x)1`$ compares the normalized number variance at bin scale $`\delta x`$ to the central-limit or Poisson expectation 1 for that quantity. The difference reflects correlations in the number distribution beyond that of a random distribution of points. The variance difference can be interpreted as the total number of correlated pairs (per-bin number variance $`\sigma _n^2`$) minus the number of Poisson-correlated pairs (self pairs $`\overline{n}`$) per particle (divided by $`\overline{n}`$). The corresponding variance difference for $`p_t`$ fluctuations is $`\mathrm{\Delta }\sigma _{p_t:n}^2(\delta x)\overline{(p_t(\delta x)n(\delta x)\widehat{p}_t)^2}/\overline{n}(\delta x)\sigma _{\widehat{p}_t}^2`$. Those per-particle quantities, defined in terms of normalized variances, are consistent in form with Pearson’s normalized covariance. A fluctuation measurement at a single scale (STAR detector acceptance) is shown in the first panel of Fig. 4 . The frequency histogram on random variable $`(p_tn\widehat{p}_t)/(\sqrt{\overline{n}}\sigma _{\widehat{p}_t})`$ is compared to a central-limit reference (green dotted curve, $`\sigma _{\widehat{p}_t}`$ is the ensemble-average single-particle variance). The $`p_t`$ variance difference $`\mathrm{\Delta }\sigma _{p_t:n}^2(\delta x)`$ defined previously compares variances of data and reference to reveal a variance excess. Questions then arise how to interpret the fluctuation measurement and how to compare it to measurements made with other detectors. Fluctuation measurements in different bin-size or scale intervals determine different regions on a common distribution representing fluctuation scale dependence, as shown in the second panel of Fig. 4 . The variance difference from the first panel corresponds to the single point at the apex of the surface in the second panel. The surface is obviously structured, but what does the structure mean? Fluctuation scale dependence is the running integral of an autocorrelation. The corresponding integral equation is a linear relation between a variance difference and an autocorrelation, including a kernel representing the binning scheme. We can express the per-particle variance difference on scales $`(\delta \eta ,\delta \varphi )`$ as 2D discrete integral $`\mathrm{\Delta }\sigma _{p_t:n}^2(mϵ_\eta ,nϵ_\varphi )=4{\displaystyle \underset{k,l=1}{\overset{m,n}{}}}ϵ_\eta ϵ_\varphi `$ $`K_{mn;kl}{\displaystyle \frac{\mathrm{\Delta }\rho (p_t:n;kϵ_\eta ,lϵ_\varphi )}{\sqrt{\rho _{ref}(n;kϵ_\eta ,lϵ_\varphi )}}},`$ (5) with kernel $`K_{mn;kl}(mk+1/2)/m(nl+1/2)/n`$ representing the 2D macrobin system. This is a Fredholm integral equation which can be inverted by standard numerical methods to obtain autocorrelation density ratio $`\mathrm{\Delta }\rho _{kl}/\sqrt{\rho _{ref,kl}}`$ as a per-particle correlation measure on difference variables $`(\eta _\mathrm{\Delta },\varphi _\mathrm{\Delta })`$ . The third panel of Fig. 4 shows the autocorrelation corresponding to the data in the second panel, with directly interpretable structure . ## 8 Derivation of the integral equation The derivation relies on coordinating the contents of macrobins, which determine the binning scale for fluctuations, and microbins, which are the basis for the discrete numerical integration of the autocorrelation. We divide the acceptance into macrobins of varying size $`\delta x`$ (scale dependence) and microbins of fixed size $`ϵ_x`$. The scale-dependent variance is an average over all bins of a particular scale within the acceptance as in Fig. 5 (first panel), where the acceptance is $`\mathrm{\Delta }x`$, the macrobin size is $`\delta x`$ and the macrobin number is $`M`$. The average over macrobins is re-expressed as an average over microbins, as shown in Fig. 5 (second panel). The average over microbins is re-arrange into a sum (over $`k`$) of different diagonals and an average over microbins on the $`k^{th}`$ diagonal. The diagonal average is the $`k^{th}`$ element of an autocorrelation histogram. The $`p_t`$ variance difference at scale $`\delta x`$ in the first line of Eq. (8) is re-expressed as a 2D sum over microbins in the second line, with $`m`$ microbins in each macrobin. Mean multiplicity $`\overline{n}(\delta x)`$ in a macrobin relates to the mean microbin multiplicity in that macrobin as $`\overline{n}(\delta x)=m\overline{n}(ϵ_x)`$, which relation is applied in the second line. The single-particle variance in the first line corresponds to self pairs which are excluded from subsequent pair sums. The single-particle variance term is consequently dropped. The third line is a rearrangement of the sum over $`(a,b)`$ into a sum over diagonal index $`k`$ and a sum over indices $`a,b`$ subject to the diagonal constraint $`ab=k`$, $`\mathrm{\Delta }\sigma _{p_t:n}^2(\delta x)`$ $``$ $`\overline{(p_t(\delta x)n(\delta x)\widehat{p}_t)^2}/\overline{n}(\delta x)\sigma _{\widehat{p}_t}^2`$ $`=`$ $`{\displaystyle \underset{a,b=1}{\overset{m}{}}}{\displaystyle \frac{\overline{\{p_t(ϵ_x)n(ϵ_x)\widehat{p}_t\}_a\{p_t(ϵ_x)n(ϵ_x)\widehat{p}_t\}_b}}{m\overline{n}(ϵ_x)}}`$ $`=`$ $`{\displaystyle \underset{k=1m}{\overset{m1}{}}}K_{m:k}{\displaystyle \frac{\overline{n}_k}{\overline{n}}}\left[{\displaystyle \frac{1}{m|k|}}{\displaystyle \underset{1a,bm}{\overset{ab=k}{}}}{\displaystyle \frac{\sqrt{\overline{n}_a\overline{n}_b}}{\overline{n}_k}}{\displaystyle \frac{\overline{\{\mathrm{}\}_a\{\mathrm{}\}_b}}{\sqrt{\overline{n}_a\overline{n}_b}}}\right]`$ $``$ $`{\displaystyle \underset{k=1m}{\overset{m1}{}}}K_{m:k}{\displaystyle \frac{\mathrm{\Delta }A_k(p_t:n;ϵ_x)}{\sqrt{A_{k,ref}(n;ϵ_x)}}}`$ $``$ $`2{\displaystyle \underset{k=1}{\overset{m}{}}}K_{m:k}{\displaystyle \frac{\mathrm{\Delta }A_k(p_t:n;ϵ_x)}{\sqrt{A_{k,ref}(n;ϵ_x)}}}2{\displaystyle \underset{k=1}{\overset{m}{}}}ϵ_xK_{m:k}{\displaystyle \frac{\mathrm{\Delta }\rho (p_t:n;kϵ_x)}{\sqrt{\rho _{ref}(n;kϵ_x)}}}.`$ (7) Factors have been introduced so that the expression in square brackets is an average along the $`k^{th}`$ diagonal of normalized microbin covariances, with weighting factor $`\sqrt{\overline{n}_a\overline{n}_b}/\overline{n}_k1`$. The fourth line identifies the square bracket as a ratio of autocorrelation histograms. Factor $`\overline{n}_k/\overline{n}1`$ has been absorbed into the definition of the autocorrelation ratio, but could be extracted as a correction factor. In the fifth line the binning system has been shifted by 1/2 bin according to Fig. 5 (third panel), and symmetry about the origin has been invoked (requiring an additional factor 2) to simplify the indexing. The histograms are finally converted to densities by including the microbin width. The last line, generalized to 2D on angle variables $`(\eta ,\varphi )`$, is Eq. (5). That integral equation provides computationally cheap $`O(n)`$ access to autocorrelations. ## 9 Inversion and regularization Eq. (5), a discrete Fredholm integral equation, is a matrix equation of the form $`𝐃=𝐓𝐈+𝐍`$, with data D, image I and statistical noise $`𝐍`$ . In principle, one could simply invert the matrix equation to obtain the image. However, $`𝐓^\mathrm{𝟏}`$ is effectively a differentiation and acts therefore as a high-pass filter in the language of electrical engineering. The $`𝐓^\mathrm{𝟏}𝐍`$ statistical noise term strongly dominates the image derived from a simple matrix inversion, and must be substantially reduced by smoothing or ‘regularization’ . The procedure involves treating the image as a matrix of free values in a $`\chi ^2`$ fit subject to Tikhonov regularization: minimize $`\chi _\alpha ^2𝐃𝐓𝐈_\alpha ^\mathrm{𝟐}+\alpha 𝐋𝐈_\alpha ^\mathrm{𝟐}`$, including Lagrange multiplier $`\alpha `$ which controls the role of local gradient operator $`𝐋`$. The first term measures the data–integrated-image mismatch, the second term measures small-wavelength noise on the image. The latter is equivalent to a compensating low-pass filter which offsets the effect of the differentiation. The resulting image is represented by $`𝐈_\alpha =𝐓_\alpha ^\mathrm{𝟏}(𝐃𝐍_\alpha )`$ which estimates true image $`𝐈`$. Regularization is illustrated in a power-spectrum context by Fig. 6 (first panel): a tradeoff between information loss and noise suppression. The basis for choosing the optimum $`\alpha `$ is illustrated in the second panel. $`𝐃𝐓𝐈_\alpha ^\mathrm{𝟐}`$ (dots) is signal loss and $`𝐋𝐈_\alpha ^\mathrm{𝟐}`$ (triangles) is residual noise. As noted, $`\alpha `$ controls a compensating low-pass filter: small values retain all signal and a large amount of noise from the differentiation, larger values reduce noise, and finally distort the signal by over smoothing. The optimum value is determined as in Fig. 6 (second panel). This example illustrates that there are clear criteria for choosing an optimum $`\alpha `$ so that negligible information is lost from the image while statistical noise is greatly attenuated. Statistical and systematic errors are determined by looping through integration and differentiation twice in the sequence $`𝐃𝐈_\alpha 𝐃_\alpha 𝐈_{\alpha \alpha }`$, including inversion, forward integration and second inversion. Difference $`𝐃𝐃_\alpha `$ estimates statistical error on the data and may itself be inverted to determine residual statistical error on image $`𝐈_\alpha `$. Difference $`𝐈_\alpha 𝐈_{\alpha \alpha }`$ estimates the smoothing error. Fig. 6 (right two panels) shows image $`𝐈_\alpha `$ and corresponding smoothing error $`𝐈_\alpha 𝐈_{\alpha \alpha }`$ for data from the Hijing Monte Carlo , dominated by jet correlations . A precision comparison of autocorrelations obtained directly by pair counting and indirectly by fluctuation inversion is presented in Fig. 7. The data were obtained from the Pythia Monte Carlo . The four panels in sequence are charge-independent direct and inverted autocorrelations and charge-dependent direct and inverted autocorrelations. The agreement between methods is excellent. ## 10 Correlation types Fig. 8 illustrates types of correlation measurements. The left panels show an autocorrelation obtained by $`p_t`$ fluctuation inversion (first panel) and a sketch of corresponding two-bin correlations (second panel) for an arbitrary bin separation. In this type of correlation, structure may occur in most or all events, but with random position on $`(\eta ,\varphi )`$. The autocorrelation of a positive-definite measure (such as $`n`$ or $`p_t`$) must be positive-definite at the origin (a variance). Elsewhere (covariances), positive autocorrelation bins indicate correlation (solid curve in second panel), negative bins indicate anticorrelation (dash-dot curve), as in the magenta areas adjacent to the positive same-side peak in the first panel. The dashed circle represents the reference. The right panels illustrate a case where localized structure with approximately fixed position is present in a minority of events. The third panel shows correlations on transverse rapidity $`y_t`$ (that distribution is not an autocorrelation). In the fourth panel distributions are sketched for two event classes. The dashed curve represents common (e.g., soft) events, the solid circle represents exceptional events (e.g., hard events which contain a detectable parton scatter) from an ensemble of p-p collisions. Positive covariances result from the exceptional events which have an additional multiplicity contribution localized on the $`y_t`$ space and (in contrast to the autocorrelation example at left) occuring at a nearly fixed position over the event ensemble. The normalized covariance density comparing sibling and mixed pairs in the third panel reveals the contribution from exceptional events (parton fragments). ## 11 Autocorrelations and conditional distributions Conventional studies of jet correlations in A-A collisions employ a leading-particle analysis (invoked when full jet reconstruction is not possible) . The goal is to estimate a parton momentum by that of the highest-$`p_t`$ (above some threshold) particle in a collision—the leading or trigger particle. The analysis utilizes two or three conditional distributions as illustrated in Fig. 9 (first two panels), where for the sake of comparison transverse momentum $`p_t`$ has been replaced by transverse rapidity $`y_t`$. The condition on $`(y_{t1},y_{t2})`$ is a rectangle representing asymmetric trigger- and associated-particle $`y_t`$ conditions. Trigger region $`(\mathrm{\Omega }_{t_{t1}},\mathrm{\Omega }_{t_{t2}})`$ is by construction displaced from the sum diagonal. Angular correlations are also defined as conditional distributions relative to the trigger-particle position in single-particle angle space $`(\eta ,\varphi )`$. The angular correlation is plotted (using pseudorapidity as an example) on conditional angle difference $`\mathrm{\Delta }\eta \eta \eta _{trigger}`$, which represents an event-wise shift of the single-particle angle origin. Alternatively, one can abandon attempts to estimate parton $`p_t`$ per se and use a technique involving no conditions on momentum, or a symmetric condition on $`(y_{t1},y_{t2})`$ with no conditions on angle variables, as illustrated in Fig. 9 (second two panels). The fragment distribution on $`(y_{t1},y_{t2})`$ can be studied in its own right. Cut conditions on $`(y_{t1},y_{t2})`$ can be defined to study corresponding changes in jet angular morphology on ($`\eta ,\varphi `$). Angular autocorrelations invoking no trigger condition and defined on symmetric difference variables such as $`\eta _\mathrm{\Delta }\eta _1\eta _2`$ access a minimum-bias parton population. Difference variable $`\eta _\mathrm{\Delta }`$ spans the diagonal axis of the 2D $`(\eta _1,\eta _2)`$ space. Variables $`\mathrm{\Delta }\eta `$ and $`\eta _\mathrm{\Delta }`$ are numerically different, support different distributions and should not be confused (the same comment applies to $`\mathrm{\Delta }\varphi `$ and $`\varphi _\mathrm{\Delta }`$). The correlation measures used in the two cases may be compared: $`{\displaystyle \frac{\mathrm{\Delta }\rho (n;\eta _\mathrm{\Delta },\varphi _\mathrm{\Delta })}{\sqrt{\rho _{ref}(n;\eta _\mathrm{\Delta },\varphi _\mathrm{\Delta })}}}vs{\displaystyle \frac{1}{N_{trig}}}{\displaystyle \frac{d^2N_{pair}}{d\mathrm{\Delta }\eta d\mathrm{\Delta }\varphi }}{\displaystyle \frac{\rho (n;\mathrm{\Delta }\eta ,\mathrm{\Delta }\varphi )}{_{\mathrm{\Omega }_{trig}}𝑑p_t\rho (n;p_{t,trig})}}.`$ (8) The latter depends directly on leading- or trigger-particle acceptance $`\mathrm{\Omega }_{trig}`$, and a reference must be provided a posteriori by defining a model function to describe the background. ## 12 Summary We have derived an integral equation which connects fluctuation scale dependence to corresponding autocorrelations. Inversion of the integral equation reveals autocorrelations which are equivalent to those from pair counting. Fluctuations are thereby interpretable in terms of underling two-particle correlations. Autocorrelations are complementary to leading-particle techniques for analysis of jet correlations. Definition of normalized variances, covariances and variance differences is tightly constrained by a number of considerations, including linearity, relating fluctuations to correlations, minimizing statistical bias and insuring that correlation structure is maximally accessible and interpretable over a broad range of contexts. This work was supported in part by the Office of Science of the U.S. DoE under grant DE-FG03-97ER41020.
warning/0506/cond-mat0506625.html
ar5iv
text
# Complete spin extraction from semiconductors near ferromagnet-semiconductor interfaces ## Abstract We show that spin polarization of electrons in nonmagnetic semiconductors near specially tailored ferromagnet-semiconductor junctions can achieve 100%. This effect is realized even at moderate spin injection coefficients of the contact when these coefficients only weakly depend on the current. The effect of complete spin extraction occurs at relatively strong electric fields and arises from a reduction of spin penetration length due to the drift of electrons from a semiconductor towards the spin-selective tunnel junction. Combining carrier spin as a new degree of freedom with the established bandgap engineering of modern devices offers exciting opportunities for new functionality and performance. This new field of semiconductor physics is referred to as semiconductor spintronics Zut ; Aw . The injection of spin-polarized electrons into nonmagnetic semiconductors (NS) is of particular interest because of the relatively large spin-coherence lifetime, $`\tau _s`$, and the promise for applications in both ultrafast low-power electronic devices Zut ; Aw ; Datta ; Hot ; OBO and in quantum information processing Aw ; QIP ; QC . The main challenge is to achieve a high spin polarization, $`P_n`$, of electrons in NS. The characteristics of the spintronic devices dramatically improve when $`P_n100\%`$. It has been concluded in all previous theoretical works on spin injection Aron ; Mark ; Rash ; Flat ; Alb ; BO ; OB that $`P_n`$ cannot exceed either the spin polarization of the carriers in the spin source or the spin injection coefficient, $`\gamma `$, of the ferromagnet-semiconductor junction Remk . This conclusion does not contradict existing experiments in which different magnetic materials such as magnetic semiconductors, half-metallic ferromagnets, and ferromagnetic metals (FM) have been used as spin sources Zut ; Aw . FM are widely used in semiconductor technology. The Curie tempeartures of these materials are usually much higher than the room temperature. The greatest value of $`P_n`$ 32%, was achieved for Fe-based junctions Jonk ; Ohno with approximately the same polarization of the source. One of the obstacles for the spin injection from FM into NS is a high and wide Schottky barrier that usually forms at the metal-semiconductor interfaces sze . The spin injection corresponds to a reverse current of the Schottky FM-S junction. This current is usually extremely small sze . Therefore, a thin heavily doped $`n^+`$-S layer between FM and S must be used to increase the current sze ; Jonk ; Alb ; OB ; BO . This layer greatly reduces the thickness of the barrier and increases its tunneling transparency. The greatest values of $`P_n`$ were found in such FM- $`n^+`$-$`n`$-S structures Jonk . Thus, the spin injection is the tunneling of spin polarized electrons from FM into NS in reverse-biased FM-S structures. Since the tunneling is a symmetric process the spin selective transport must also occur in the forward-biased junctions when electrons are emitted from NS into FM BO . In these junctions the electrons with a certain spin projection can be efficiently extracted from NS while the opposite spin electrons will accumulate in NS near FM-S interface BO . Spin extraction from NS was predicted by I. Zutic et al. Zut1 for forward-biased p-n junctions containing a magnetic semiconductor and was experimentally found in forward-biased MnAs/GaAs Schottky junction Step . However the predicted and observed values of $`P_n`$ were rather small. In this letter we demonstrate a possibility for achieving complete spin polarization of electrons in NS near forward-biased FM-S junctions with moderate spin injection coefficient, $`\gamma `$. The effect is based on spin extraction and nonlinear dependence of the nonequilibrium spin density on the electric field. We consider a FM-$`n^+`$-$`n`$-S structure containing a heavily doped degenerate $`n^+`$-S layer, Fig. 1. We use a standard assumption of spin injectionAron ; Mark ; Rash ; Flat ; Alb that $`\gamma `$ of the FM-$`n^+`$-S contact only weakly depends on the total current $`J`$ due to a high density of degenerate electrons in the $`n^+`$-S layer (see below). In the forward-biased structure unpolarized electrons drift from the bulk of NS to the contact. Because of the spin selectivity of the contact the electrons with spin $`\sigma =`$ (up-electrons) at $`\gamma >0`$ are extracted from NS, i.e. $`\delta n_l=(n_ln_s/2)<0`$, and electrons with spin $`\sigma =`$ (down-electrons) are accumulated, i.e. $`\delta n_{}=(n_{}n_s/2)>0`$, near the contact BO . Here $`n_s`$, $`n_l`$ and $`n_l`$ are the equilibrium electron density in NS and densities of up-and down-electrons, respectively, at the boundary between the $`n^+`$-S layer and high-resistant NS region ($`x=l`$ in Fig. 1(a)). The quantity $`\left|\delta n_l\right|`$ increases with the electric field, $`E`$ BO . In sufficiently strong fields, the drift efficiently compresses the spin polarized electrons to the boundary. As a result Flat ; BO , the spin penetration length $`L`$ decreases with the current $`J`$ \[cf. white and red curves in Fig. 1(a)\]. Note, that due to $`\delta n_{}=\delta n_{}`$, the diffusion flow of up-electrons is directed along the electron drift while the diffusion flow of down-electrons is in the opposite direction, Fig. 1(a). The superlinear increase of the spin diffusion flows with $`J`$ can be compensated only by an increase of the spin density $`n_{}`$ up to $`n_s`$ and a decrease of $`n_{}`$ down to zero. In other words, spin polarization of the electrons in NS near FM- $`n^+`$-S contact $`\left|P_{nl}\right|=\left|\delta n_l\delta n_l\right|/n_s=2\left|\delta n_l\right|/n_s`$ can reach 100% when the current is sufficiently large. Let us consider for simplicity the case when the diffusion constant and mobility of up- and down-electrons are the same constants: $`D_{}=D_{}=D`$ and $`\mu _{}=`$ $`\mu _{}=`$ $`\mu `$. This standard assumption Aron ; Mark ; Rash ; Flat ; Alb is valid for nondegenerate NS (the peculiarities of degenerate NS are discussed below). In this case the currents of up- and down electrons with $`\sigma =,`$ are given by the equations Aron ; Flat ; OB ; BO $`J_\sigma `$ $`=`$ $`q\mu n_\sigma E+qD{\displaystyle \frac{d\delta n_\sigma }{dx}},`$ (1) $`dJ_{}/dx`$ $`=`$ $`q(n_{}n_{})/2\tau _s\text{ },`$ (2) where $`q`$ is the magnitude of the elementary charge. It follows from conditions of the continuity of the total current and electroneutrality that $`J(x)=J_{}+J_{}=\mathrm{const},`$ and $`n(x)=n_{}+n_{}=n_s=\mathrm{const}`$. This means that $`E(x)=J/q\mu n_s=\mathrm{const}`$ and $`\delta n_{}(x)=\delta n_{}(x)`$. Then the solution of Eqs. (1)-(2) reads Aron ; Flat ; OB ; BO $`\delta n_{}(x)`$ $`=`$ $`P_{nl}{\displaystyle \frac{n_s}{2}}\mathrm{exp}[(xl)/L],`$ (3) $`\text{where }L`$ $`=`$ $`(1/2)\left(\sqrt{4L_s^2+L_E^2}\pm L_E\right)`$ (4) where $`P_{nl}=P_n(l)=2\delta n_l/n_s`$ is the spin polarization of the up-electrons at $`x=l`$ (Fig. 1), $`L_s=\sqrt{D\tau _s}`$ and $`L_E=\mu \tau _s\left|E\right|=`$ $`L_s\left|J\right|/J_s`$ are the spin diffusion and drift lengths, respectively, and $`J_s=qn_SD/L_s`$. The signs $`\pm `$ correspond to the reversed, $`J<0`$, and forward biases, $`J>0`$, respectively. From Eqs. (1)-(3) we find that the currents at $`x=l`$ are $$J_{l,l}=\frac{J}{2}\pm J\frac{\delta n_l}{n_s}qD\frac{\delta n_l}{L}=\frac{J}{2}\frac{J_s}{2}\frac{L}{L_s}P_{nl}.$$ (5) It follows from Eq. (5) that the electron spin polarizations, $`P_{nl}=2\delta n_{}/n_s`$, and the spin injection coefficient, $`\gamma _l=(J_lJ_l)/J`$, near the boundary are related by the equation $$P_{nl}=\gamma _l\frac{JL_s}{J_sL}=\frac{2J\gamma _l}{\sqrt{(2J_s)^2+J^2}\pm \left|J\right|}$$ (6) Thus, we see that for the case of the spin injection (reversed bias, sign +) $`\left|P_{nl}\right|<\left|\gamma _l\right|`$ in accordance with previous works Aron ; Flat ; OB . Another situation is realized in *the forward-biased FM-S junctions*, sign $``$ in Eq. (6). Here the spin penetration depth $`L`$ (4) decreases with the current $`J`$ and according to (6) $`\left|P_{nl}\right|`$ *approaches 1 (100%)* when $`J`$ $``$ $`J_tJ_s(\left|\gamma _l\right|+\gamma _l^2)^{1/2}`$ (7) $`\text{and }L`$ $``$ $`L_tL_s\sqrt{\left|\gamma _l\right|/(1+|\gamma _l|).}`$ (8) In degenerate NS the diffusion constants depend on electron densities: $`D_\sigma /\mu _\sigma =(D/\mu )(2n_\sigma /n_s)^{2/3}`$ at low temperatures T$`\mu `$. In this case we can find $`E`$ from Eqs. (1) and $`J=J_{}(x)+J_{}(x)`$. Then, substituting $`E`$ into Eq. (1), we obtain $`J_{}`$. Using this $`J_{}`$ and Eq. (2) we find a diffusion-drift equation for $`\delta n_{}(x)`$ with a bi-spin diffusion constant, $`D(P_n)=(D/2)\left(1P_n^2\right)^{2/3}\left[(1+P_n)^{1/3}+(1P_n)^{1/3}\right]`$, which depends on $`P_n=2\delta n_{}/n_s`$. One can see that $`D(P_n)0`$ when $`\left|P_n\right|1`$. It means that the effective spin diffusion length $`L_s(P_n)=[D(P_n)\tau _s]^{1/2}`$ decreases with the current because $`\left|P_{nl}\right|1`$ near $`x=l`$. Thus, an additional mechanism of a decrease of the spin penetration length $`L`$ with current $`J`$ occurs in a degenerate NS. As a result, the decay of $`P_n(x)`$ is sharper, particularly near $`x=l`$, as shown in the inset in Fig. 2. Therefore, in degenerate NS the condition of complete spin extraction, $`\left|P_{nl}\right|=1`$, can be reached at lower threshold currents and greater spin lengths as compared with those given by (7) and (8) for nondegenerate NS. For instance, numerical analysis shows that the threshold values $`J_t=1.3J_s`$ and $`L_t=0.37L_s`$ at $`\gamma _l=0.3`$ for $`D_\sigma /\mu _\sigma =(D/\mu )(2n_\sigma /n_s)^{2/3}`$while $`J_t=1.6J_s`$ and $`L_t=0.48L_s`$ for the case $`D_\sigma /\mu _\sigma =const`$. The effect of complete spin extraction from a degenerate NS can be illustrated based on spatial and current dependences of quasi-Fermi levels $`F_{}`$ and $`F_{}`$ for up- and down-electrons, respectively (Fig. 2). Indeed, due to the spin extraction the difference between $`F_{}`$ and $`F_{}`$ near the FM-$`n^+`$\- S contact increases with the current. Therefore, the value $`F_{}`$ can reach the bottom of the conduction band $`E_c`$ in NS at $`x=l`$ (Fig.2) at the current $`J=J_t`$. This implies that $`\mathrm{\Delta }F_{}=F_{}F=\mu _s`$ at this point and $`n_l(FE_c+\mathrm{\Delta }F_{})0`$, $`n_l=(n_sn_l)n_s`$, i.e. $`\left|P_{nl}\right|1`$. Here $`\mu _s=FE_c`$ and $`F`$ are the Fermi energy and the equilibrium Fermi level of electrons in NS, respectively. In reality, however, our theory, which is based on the consideration of two nonequilibrium ensembles of the up- and down-electrons, becomes invalid when $`n_l0`$. Our approach is justified only when the time of electron-electron collisions within each of these systems is much less than $`\tau _s`$. Moreover, at large currents $`J>J_t`$ the value of $`\left|P_{nl}\right|=2\left|\delta n_l\right|/n_s=\left|2n_ln_s\right|/n_s`$ becomes greater than 1 (see e.g. (6)), i.e. spin density $`n_l`$ at $`x=l`$ exceeds the equilibrium electron density, $`n_s`$. Therefore, the condition of local electroneutrality $`n_l+`$ $`n_l=n_s`$ is violated and a space charge arises near $`x=l`$ in Fig.1. This charge will change the field $`E(x)`$ and the total electron density in the vicinity of $`x=l`$. The complete set of equations consists of Eqs. (1)-(2), $`J=J_{}(x)+J_{}(x)=\mathrm{const}`$, the and Poisson’s equation: $`\epsilon \epsilon _0dE/dx=\rho `$, where $`\rho =q(n_sn_{}+`$ $`n_{})`$ and $`\epsilon \epsilon _0`$ is the permittivity of the NS. Our calculations for the case of $`\gamma _l=const`$ show that, as expected, the characteristic scale of the nonuniform-field region is determined by a relatively short screening length and the value of $`\left|P_{nl}\right|`$ in the degenerate NS is close to $`1`$ near $`x=l`$ at $`JJ_t`$. One can see from (7) and (8) that the spin injection coefficient of FM-$`n^+`$-S contact, $`\gamma _l`$, determines the threshold current, $`J_t`$, and spin penetration depth, $`L_t`$. However our main finding that $`\left|P_{nl}\right|1`$ at $`JJ_t`$ remains vlid at any reasonable value of $`\gamma _l`$. The only required condition is a relatively weak dependence of $`\gamma _l`$ on $`J`$ (see Remark ). This can be realized in a FM-$`n^+`$-S junction containing a heavily doped $`n^+`$-S layer. The donor concentration, $`N_d^+`$, and thickness, $`l`$, of this layer must satisfy the following conditions: $`l3w`$ and $`N_d^+w^2q^22\epsilon \epsilon _0\mathrm{\Delta }`$, where $`\mathrm{\Delta }`$ and $`w`$ are the height and width of the depletion Schottky layer, Fig. 1. The electron gas has to be highly degenerate in a certain part of the $`n^+`$-S layer contiguous $`n`$-S layer. The transition between the $`n^+`$-S and $`n`$-S layers should have a discontinuous jump $`\mathrm{\Delta }_0=(E_cE_c^+)`$ shown in Fig.1(b). This is realized when the $`n^+`$-S layer has a narrower energy bandgap than that of the $`n`$-S region. A similar diagram can also be realized when $`n^+`$-S and $`n`$-S regions are made of the same semiconductor, but an additional, acceptor-doped, ultrathin layer is formed between the $`n^+`$-S and $`n`$-S regions. The acceptor concentration $`N_a`$ and thickness $`l_a`$ of this layer have to satisfy the conditions: $`N_al_a^2q^22\epsilon \epsilon _0\mathrm{\Delta }_0`$ and $`l_al`$. To demonstrate the weak dependence of $`\gamma _l`$ upon the current $`J`$ through FM-$`n^+`$-S junction we use the common assumption that the electron energy $`E`$, spin $`\sigma ,`$ and the lateral component $`\stackrel{}{k}_{}`$ of the wave vector $`\stackrel{}{k}`$ are conserved during tunneling. Then the current density of electrons with spin $`\sigma =,`$ tunneling through the Schottky barrier, i.e. between the points $`x=w`$ and $`x=0`$ in Fig. 1, can be expressed as OB ; BO : $$J_{\sigma w}=\frac{q}{h}𝑑E[f(EF_{\sigma w}^+)f(EF)]\frac{d^2k_{}}{(2\pi )^2}T_{k\sigma },$$ (9) where $`f(EF)`$ is the Fermi function, $`F`$ the Fermi level in FM, $`F_{\sigma w}^+`$ quasi-Fermi levels up- and down-electrons in $`n^+`$\- S layer near the FM-S interface ($`x=w`$ in Fig.1), and $`T_{k\sigma }`$ is the transmission probability. We also assume that the temperature $`T\mu _s^+/k_B`$, and $`\mu _s^+=(FE_{c0}^+)`$ is the Fermi energy of degenerate equilibrium electrons of the $`n^+`$-S layer. In this case the nonequilibrium density of the electrons with spin $`\sigma `$ at $`x=w`$ reads $$n_{\sigma w}^+=\frac{n^+}{2(\mu _s^+)^{3/2}}(F_{\sigma w}^+E_{c0}^+qV)^{3/2}=\frac{n^+}{2}\left[1+\frac{\mathrm{\Delta }F_{\sigma w}^+}{\mu _s^+}\right]^{3/2},$$ (10) where $`n^+`$ is the equilibrium electron density at $`x=w`$$`E_{c0}^+`$ is the bottom of conduction band in the $`n^+`$-S region in equilibrium, $`V`$ is the bias voltage, and $`\mathrm{\Delta }F_{\sigma w}^+=(F_{\sigma w}^+FqV)`$. Using the approximate expression for $`T_{k\sigma }`$ BO ; OB , and Eqs. (9) - (10) at $`T\mu _s/k_B`$, $`\left|qV\right|<\mu _s^+`$ and $`w3l_0`$ we obtain $$J_{\sigma w}=j_0d_\sigma T_0(\mu _s^+)^{5/2}\left[\left(\mu _s^++\mathrm{\Delta }F_{\sigma w}^+\right)^{5/2}\left(\mu _s^+qV\right)^{5/2}\right]$$ (11) where $`j_0=qn_s^+v_F\alpha _0`$, $`\alpha _00.96(\kappa _0l)^{1/3}1`$ and $`T_0=\mathrm{exp}\left[\eta w\frac{(\mathrm{\Delta }\mu _s^+qV)^{1/2}}{l_0\mathrm{\Delta }^{1/2}}\right]`$ and $`d_\sigma =v_Fv_{\sigma 0}/(v_{t0}^2+v_{\sigma 0}^2)`$ is the tunneling transparency and the spin selection factor of FM-$`n^+`$-S contact; $`\eta 4/3`$, $`l_0=(\mathrm{}^2/2m_{}\mathrm{\Delta })^{1/2}`$ is a tunneling length, $`v_{t0}=\sqrt{2(\mathrm{\Delta }qV)/m_{}}`$, $`v_{\sigma 0}=v_\sigma (F+qV)`$ and $`v_F=\sqrt{3\mu _s^+/m_{}}`$ are velocities of electrons with spin $`\sigma `$ and the energies $`F+qV`$ and $`\mu _s^+`$ in FM and $`n^+`$-S regioons, respectively, and $`m_{}`$ effective mass of electrons in $`n^+`$-S layer. Let us consider the case when the thickness of the $`n^+`$-S layer $`lL_s^+`$, but $`l3w`$. Here $`L_s^+=\sqrt{D^+\tau _s^+}`$ is the spin diffusion length in the $`n^+`$-S layer. Due to the condition $`lL_s^+`$ the quasi-Fermi levels, $`F_{}`$ and $`F_{}`$ and the spin currents change very weakly in the $`n^+`$-S layer (Fig. 2). Therefore we can put $`J_{\sigma w}J_{\sigma l}`$ and $`\gamma _w\gamma _l`$. We noticed above that in degenerate $`n^+`$-S $`\left|\mathrm{\Delta }F_{\sigma w}^+\right|\mu _s=(E_cF)`$ at $`x=w`$ when $`JJ_t`$. Due to $`n_s^+n_s`$ the value $`\mu _s^+(n_s^+)^{2/3}\mu _s`$, therefore we can neglect $`\left|\mathrm{\Delta }F_{\sigma w}^+\right|`$ in Eq.(11) in comparison with $`\mu _s^+`$ when $`qV\mu _s^+`$ at $`JJ_t`$. In this case we find that the spin injection coefficient, $`\gamma _w=(J_wJ_w)/J,`$ and the total current of the FM-S junction are equal $`\gamma _0`$ $`=`$ $`(d_{}d_{})/(d_{}+d_{})`$ (12) $`J`$ $`=`$ $`J_0T_0\left[1\left(1qV/\mu _s^+\right)^{5/2}\right].`$ (13) Here $`J_0=(d_{}+d_{})j_0`$ and $`\gamma _0`$ depend weakly on $`V`$ and $`J`$ ($`\gamma _0`$ can increase with $`V`$ BO ). We note that $`J_0n_s^+=N_d^+`$ while $`J_tn_s=N_d`$, and, therefore $`J_0J_t`$. We see that $`\gamma _l\gamma _w=\gamma _0`$ in the forward-biased FM-$`n^+`$-$`n`$-S structures when $`L_s^+>l3l_D`$, $`J_0J_t`$, and $`J_0T_0J_t`$ at $`qV\mu _s^+`$. In other words we suppose that Rashba’s condition Rash is valid for the FM-$`n^+`$-S junction and therefore the spin injection coefficient $`\gamma _l`$ only weakly depends on the current at $`JJ_t`$. In this case, as we have shown above, the spin polarization of electrons in degenerate $`n`$-S region near the $`n^+`$-S layer, $`P_{nl}100\%`$ as $`JJ_t`$. In real ferromagnets the situation is much more complex. In FMs there are spin-polarized heavy d-electrons and nonpolarized light s-electrons with very involved energy spectrum. Nonetheless our conclusion about the weak dependence of the spin injection coefficient $`\gamma _l`$ on the current remains valid for any complex spectrum. This conclusion is based on the fact that the perturbations of the quasi-Fermi levels in $`n^+`$-S layer are small: $`\mathrm{\Delta }F_{\sigma 0}^+qV\mu _s^+`$. The latter inequality follows from a very large mismatch of the carrier concentrations in the heavily doped $`n^+`$-S layer and NS region with higher resistivity: $`n_s/n_s^+=N_d/N_d^+1`$. In conclusion, we emphasize that we have demonstrated a possibility of achieving 100% spin polarization in NS via electrical spin extraction, using FM-S contacts with moderate spin injection coefficients that weakly depend on the current. The highly spin-polarized electrons, according to the results of Ref. Kaw , can be efficiently utilized to polarize nuclear spins in semiconductors. They can also be used to spin polarize electrons on impurity centers or in quantum dots located near the FM-S interface. These effects are important for spin-based quantum information processing Aw ; QIP ; QC . The considered FM-$`n^+`$-$`n`$-S heterostructures and FM-$`n^+`$-S contacts can be used as very efficient spin polarizers or spin filters in most of the spin devices proposed to date Zut ; Aw ; Datta ; Hot ; OBO . In particular, such devices as spin-based high-frequency spin-transistors, square law detectors, frequency multipliers, magnetic sensors OBO , spin-light emitting diodes (spin-LEDs) Jonk ; BORad , and spin-resonant tunneling diodes (spin-RTDs) Pet can be modified to significantly enhance their performance.
warning/0506/cond-mat0506664.html
ar5iv
text
# 1.1 Thermodynamic functionals for non equilibrium systems ## 1.1 Thermodynamic functionals for non equilibrium systems In equilibrium statistical mechanics there is a well defined relationship, established by Boltzmann, between the probability of a state and its entropy. This fact was exploited by Einstein to study thermodynamic fluctuations. So far it does not exist a theory of irreversible processes of the same generality as equilibrium statistical mechanics and presumably it cannot exist. While in equilibrium the Gibbs distribution provides all the information and no equation of motion has to be solved, the dynamics plays the major role in non equilibrium. When we are out of equilibrium, for example in a stationary state of a system in contact with two reservoirs, even if the system is in a local equilibrium state so that it is possible to define the local thermodynamic variables e.g. density or magnetization, it is not completely clear how to define the thermodynamic potentials like the entropy or the free energy. One possibility, adopting the Boltzmann–Einstein point of view, is to use fluctuation theory to define their non equilibrium analogs. In fact in this way extensive functionals can be obtained although not necessarily simply additive due to the presence of long range correlations which seem to be a rather generic feature of non equilibrium systems. This possibility has been pursued in recent years leading to a considerable number of interesting results. One can recognize two main lines. The first, as well exemplified by the work of Derrida, Lebowitz and Speer , consists in exact calculations in specific models of stochastic lattice gases. The second is based on a macroscopic dynamical approach for Markovian microscopic evolutions, of which stochastic lattice gases are a main example, that leads to some general variational principles. We introduced this approach in and developed it in . Both approaches have been very effective and of course give the same results when a comparison is possible. Let us recall the Boltzmann–Einstein theory of equilibrium thermodynamic fluctuations, as described for example in . The main principle is that the probability of a fluctuation in a macroscopic region of fixed volume $`V`$ is $$P\mathrm{exp}\{V\mathrm{\Delta }S/k\}$$ (1.1) where $`\mathrm{\Delta }S`$ is the variation of the specific entropy calculated along a reversible transformation creating the fluctuation and $`k`$ is the Boltzmann constant. Eq. (1.1) was derived by Einstein simply by inverting the Boltzmann relationship between entropy and probability. He considered (1.1) as a phenomenological definition of the probability of a state. Einstein theory refers to fluctuations for equilibrium states, that is for systems isolated or in contact with reservoirs characterized by the same chemical potentials. When in contact with reservoirs $`\mathrm{\Delta }S`$ is the variation of the total entropy (system + reservoirs) which for fluctuations of constant volume and temperature is equal to $`\mathrm{\Delta }F/T`$, that is minus the variation of the free energy of the system divided by the temperature. We consider a stationary nonequilibrium state (SNS), namely, due to external fields and/or different chemical potentials at the boundaries, there is a flow of physical quantities, such as heat, electric charge, chemical substances, across the system. To start with it is not always clear that a closed macroscopic dynamical description is possible. If the system can be described by a hydrodynamic equation, a fact which can be rigorously established in stochastic lattice gases, a reasonable goal is to find an explicit connection between the thermodynamic potentials and the dynamical macroscopic properties like transport coefficients. As we discussed in , the study of large fluctuations provides such a connection. It leads in fact to a dynamical theory of the free energy which is shown to satisfy a Hamilton-Jacobi equation in infinitely many variables requiring as input the transport coefficients. In the case of homogeneous equilibrium states the solution of the Hamilton-Jacobi equation is easily found, and the equilibrium free energy is recovered together with the well known fluctuation-dissipation relationship, widely used in the physical and physical-chemical literature. On the other hand in SNS the Hamilton-Jacobi equation is hard to solve. There are few one-dimensional models where it reduces to a non linear ordinary differential equation which, even if it cannot be solved explicitly, leads to the important conclusion that the non equilibrium free energy is a non local functional of the thermodynamic variables. This implies that correlations over macroscopic scales are present. The existence of long range correlations is probably a generic feature of SNS and more generally of situations where the dynamics is not invariant under time reversal . As a consequence if we divide a system into subsystems the free energy is not necessarily simply additive. Besides the definition of thermodynamic potentials, in a dynamical setting a typical question one may ask is the following: what is the most probable trajectory followed by the system in the spontaneous emergence of a fluctuation or in its relaxation to an equilibrium or a stationary state? To answer this question one first derives a generalization of the Boltzmann-Einstein formula from which the most probable trajectory can be calculated by solving a variational principle. The free energy is then related to the logarithm of the probability of such a trajectory and satisfies the Hamilton-Jacobi equation associated to this variational principle. For equilibrium states and small fluctuations an answer to this type of questions was given by Onsager and Machlup in 1953 . The Onsager-Machlup theory gives the following result under the assumption of time reversibility of the microscopic dynamics: the most probable creation and relaxation trajectories of a fluctuation are one the time reversal of the other. As we show in , for SNS the Onsager-Machlup relationship has to be modified in the following way: the spontaneous emergence of a macroscopic fluctuation takes place most likely following a trajectory which can be characterized in terms of the time reversed process. ## 1.2 Macroscopic dynamics and large fluctuations We consider many-body systems in the limit of infinitely many degrees of freedom. Microscopically we assume that the evolution is described by a Markov process $`X_\tau `$ which represents the state of the system at time $`\tau `$. This hypothesis probably is not so restrictive because also the dynamics of Hamiltonian systems interacting with thermostats finally is reduced to the analysis of a Markov process, see e.g. . To be more precise $`X_\tau `$ represents the set of variables necessary to specify the state of the microscopic constituents interacting among themselves and with the reservoirs. The SNS is described by a stationary, i.e. invariant with respect to time shifts, probability distribution $`P_{st}`$ over the trajectories of $`X_\tau `$. We denote by $`\mu `$ the invariant measure of the process $`X_\tau `$. The measure $`\mu `$ is a probability on the configuration space and for each fixed time $`\tau `$ we have $`P_{st}(X_\tau =\omega )=\mu (\omega )`$. We assume that the system admits a macroscopic description in terms of density fields which are the local thermodynamic variables $`\rho _i`$. The usual macroscopic interpretation of Markovianity is that the time derivatives of the thermodynamic variables $`\dot{\rho }_i`$ at a given instant of time depend only on the $`\rho _i`$’s and the affinities (thermodynamic forces) $`\frac{F}{\rho _i}`$ at the same instant, recall that $`F`$ is the free energy. As we discussed in , for non equilibrium systems, the affinities, defined as the derivative of the non equilibrium free energy, do not determine the macroscopic evolution of the variables $`\rho _i`$. There is an additional non dissipative term which however does not modify the rate of approach to the stationary state. For simplicity of notation we assume that there is only one thermodynamic variable $`\rho `$ e.g. the local density. For conservative systems the evolution of the field $`\rho =\rho (t,u)`$, where $`t`$ and $`u`$ are the macroscopic time and space coordinates, is then given by the continuity equation $$_t\rho =\left[\frac{1}{2}D(\rho )\rho \chi (\rho )E\right]=J(\rho )$$ (1.2) where $`D(\rho )`$ is the diffusion matrix, $`\chi (\rho )`$ the mobility and $`E`$ the external field. Finally the interaction with the reservoirs appears as boundary conditions to be imposed on solutions of (1.2). We shall denote by $`\overline{\rho }=\overline{\rho }(u)`$ the unique stationary solution of (1.2), i.e. $`\overline{\rho }`$ is the typical density profile for the SNS. This equation derives from the underlying microscopic dynamics through an appropriate scaling limit in which the microscopic time and space coordinates $`\tau ,x`$ are rescaled diffusively: $`t=\tau /N^2`$, $`u=x/N`$ where $`N`$ is the linear size of the system so that the number of degrees of freedom is proportional to $`N^d`$. The hydrodynamic equation (1.2) represents a law of large numbers with respect to the probability measure $`P_{st}`$ conditioned on an initial state $`X_0`$. This conditional probability will be denoted by $`P_{X_0}`$. The initial conditions for (1.2) are determined by $`X_0`$. Of course many microscopic configurations give rise to the same value of $`\rho (0,u)`$. In general $`\rho =\rho (t,u)`$ is the limit of the local density $`\pi _N(X_\tau )`$. The free energy $`F(\rho )`$, defined as a functional of the density profile $`\rho =\rho (u)`$, gives the asymptotic probability of density fluctuations for the invariant measure $`\mu `$. More precisely $$\mu \left(\pi _N(X)\rho \right)\mathrm{exp}\left\{N^dF(\rho )\right\}$$ (1.3) where $`d`$ is the dimensionality of the system, $`\pi _N(X)\rho `$ means closeness in some metric and $``$ denotes logarithmic equivalence as $`N\mathrm{}`$. In the above formula we omitted the dependence on the temperature since it does not play any role in our analysis; we also normalized $`F`$ so that $`F(\overline{\rho })=0`$. In the same way, the behavior of space time fluctuations can be described as follows. The probability that the evolution of the random variable $`\pi _N(X_\tau )`$ deviates from the solution of the hydrodynamic equation and is close to some trajectory $`\widehat{\rho }(t)`$ is exponentially small and of the form $$P_{st}\left(\pi _N(X_{N^2t})\widehat{\rho }(t),t[t_1,t_2]\right)\mathrm{exp}\left\{N^d\left[F(\widehat{\rho }(t_1))+_{[t_1,t_2]}(\widehat{\rho })\right]\right\}$$ (1.4) where $`(\widehat{\rho })`$ is a functional which vanishes if $`\widehat{\rho }(t)`$ is a solution of (1.2) and $`F(\widehat{\rho }(t_1))`$ is the free energy cost to produce the initial density profile $`\widehat{\rho }(t_1)`$. Therefore $`(\widehat{\rho })`$ represents the extra cost necessary to follow the trajectory $`\widehat{\rho }(t)`$ in the time interval $`[t_1,t_2]`$. Equation (1.4) is a dynamical generalization of the Boltzmann-Einstein formula, we shall refer to it as the *dynamical large deviation principle* with dynamical*rate functional* $``$. For stochastic lattice gases, as shown in , the functional $``$ can be calculated explicitly. To determine the most probable trajectory followed by the system in the spontaneous creation of a fluctuation, we consider the following physical situation. The system is macroscopically in the stationary state $`\overline{\rho }`$ at $`t=\mathrm{}`$ but at $`t=0`$ we find it in the state $`\rho `$. According to (1.4) the most probable trajectory is the one that minimizes $``$ among all trajectories $`\widehat{\rho }(t)`$ connecting $`\overline{\rho }`$ to $`\rho `$ in the time interval $`[\mathrm{},0]`$. As shown in this minimization problem gives the non equilibrium free energy, i.e. $$F(\rho )=\underset{\widehat{\rho }}{inf}_{[\mathrm{},0]}(\widehat{\rho })$$ (1.5) To this variational principle it is naturally associated a Hamilton-Jacobi equation which plays a crucial role in the analysis developed in . We emphasize that the functional $``$, hence the corresponding Hamilton-Jacobi equation for $`F`$, is determined by the macroscopic transport coefficients $`D(\rho )`$ and $`\chi (\rho )`$, which are experimentally accessible, see e.g. . We can thus regard (1.5) as a far reaching generalization of the fluctuation-dissipation theorem since it allows to express a static quantity like the free energy in terms of the dynamical macroscopic features of the system. ## 1.3 Current fluctuation and related thermodynamic functionals Beside the density, a very important observable is the current flux . This quantity gives informations that cannot be recovered from the density because from a density trajectory we can determine the current trajectory only up to a divergence free vector field. We emphasize that this is due to the loss of information in the passage from the microscopic level to the macroscopic one. In the previous paper we have introduced a Boltzmann–Einstein type formula for current fluctuations. This formula shows that the asymptotic probability, as the number of degrees of freedom increases, of observing a current fluctuation $`j`$ on a space–time domain $`[0,T]\times \mathrm{\Lambda }`$ can be described by a rate functional $`_{[0,T]}(j)`$. In the present paper we develop the approach introduced in and illustrate some relevant applications. To discuss the current fluctuations, we introduce a vector-valued observable $`𝒥^N(\{X_\sigma `$, $`0\sigma \tau \})`$ of the trajectory $`X`$ which measures the local net flow of particles. As for the density, for stochastic lattice gases, we shall be able to derive a dynamical large deviations principle for the current. Recall that $`P_{X_0}`$ stands for the probability $`P_{st}`$ conditioned on the initial state $`X_0`$. Given a vector field $`j:[0,T]\times \mathrm{\Lambda }^d`$, we have $$P_{X_0}\left(𝒥^N(X)j(t,u)\right)\mathrm{exp}\left\{N^d_{[0,T]}(j)\right\}$$ (1.6) where the rate functional is $$_{[0,T]}(j)=\frac{1}{2}_0^T𝑑t[jJ(\rho )],\chi (\rho )^1[jJ(\rho )]$$ (1.7) in which we recall that $$J(\rho )=\frac{1}{2}D(\rho )\rho +\chi (\rho )E.$$ Moreover, $`\rho =\rho (t,u)`$ is obtained by solving the continuity equation $`_t\rho +j=0`$ with the initial condition $`\rho (0)=\rho _0`$ associated to $`X_0`$. The rate functional vanishes if $`j=J(\rho )`$, so that $`\rho `$ solves (1.2). This is the law of large numbers for the observable $`𝒥^N`$. Note that equation (1.7) can be interpreted, in analogy to the classical Ohm’s law, as the total energy dissipated in the time interval $`[0,T]`$ by the extra current $`jJ(\rho )`$. The functional $``$ describes the fluctuation properties of the current, the density and all observables related to them, as proved in Section 1.3. Among the many problems we can discuss within this theory, we study the fluctuations of the time average of the current $`𝒥^N`$ over a large time interval. This is the question addressed in in one space dimension by postulating an “additivity principle” which relates the fluctuation of the time averaged current in the whole system to the fluctuations in subsystems. We show that the probability of observing a given divergence free time average fluctuation $`J`$ can be described by a functional $`\mathrm{\Phi }(J)`$ which we characterize, in any dimension, in terms of a variational problem for the functional $`_{[0,T]}`$ $$\mathrm{\Phi }(J)=\underset{T\mathrm{}}{lim}\underset{j}{inf}\frac{1}{T}_{[0,T]}(j),$$ (1.8) where the infimum is carried over all paths $`j=j(t,u)`$ having time average $`J`$. The static additivity principle postulated in gives the correct answer only under additional hypotheses which are not always satisfied. Let us denote by $`U`$ the functional obtained by restricting the infimum in (1.8) to divergence free current paths $`j`$, i.e. $$U(J)=\underset{\rho }{inf}\frac{1}{2}[JJ(\rho )],\chi (\rho )^1[JJ(\rho )]$$ (1.9) where the infimum is carried out over all the density profiles $`\rho =\rho (u)`$ satisfying the appropriate boundary conditions. From (1.8) and (1.9) it follows that $`\mathrm{\Phi }U`$. In one space dimension the functional $`U`$ is the one introduced in . There are cases in which $`\mathrm{\Phi }=U`$ and in Subsection 6.1 below we give sufficient conditions on the transport coefficients $`D`$, $`\chi `$ for the coincidence of $`\mathrm{\Phi }`$ and $`U`$. On the other hand, while $`\mathrm{\Phi }`$ is always convex the functional $`U`$ may be non convex. In such a case $`U(J)`$ underestimates the probability of the fluctuation $`J`$. In we interpreted the lack of convexity of $`U`$, and more generally the strict inequality $`\mathrm{\Phi }<U`$, as a dynamical phase transition. In the present paper we investigate in more detail the occurence of this phenomenon. Let us denote by $`U^{}`$ the convex envelope of $`U`$; then $`\mathrm{\Phi }U^{}`$ and in Subsection 6.3 we give an example where $`U^{}<U`$. We shall also consider the fluctuation of the time averaged current with periodic boundary conditions. In Subsection 6.2 we discuss the behavior of $`U`$ and $`\mathrm{\Phi }`$ under appropriate conditions on the transport coefficient and the external field. In particular we show that for the Kipnis–Marchioro-Presutti (KMP) model , which is defined by a harmonic chain with random exchange of energy between neighboring oscillators, we have $`U(J)=(1/2)J^2/\chi (m)=(1/2)J^2/m^2`$, where $`m`$ is the (conserved) total energy. In addition we show, for $`J`$ large enough, $`\mathrm{\Phi }(J)<U(J)`$. This inequality is obtained by constructing a suitable travelling wave current path whose cost is less than $`U(J)`$. We mention that, by using the space-time approach introduced in , the possibility of taking advantage of travelling waves has been first envisaged by Bodineau and Derrida for the periodic simple exclusion process with external field. We refer to the discussion in Subsection 6.2 for a comparison between KMP and simple exclusion models. We study also the behavior of $``$ and $`\mathrm{\Phi }`$ under time reversal and derive a fluctuation relationship akin to the Gallavotti-Cohen theorem for the entropy production . In fact, we prove, in the present context of lattice gases, that the anti-symmetric part of $`\mathrm{\Phi }`$ is equal to the power produced by the external field and the reservoirs independently of the details of the model. From this relationship we derive a macroscopic version the fluctuation theorem for the entropy production. 2. Microscopic model As the basic microscopic model we consider a stochastic lattice gas with a weak external field and particle reservoirs at the boundary. The process can be informally described as follows. We consider particles evolving on a finite domain. At each site, independently from the others, particles wait exponential times at the end of which one of them jumps to a neighboring site. Superimposed to this dynamics, at the boundary particles are created and annihilated at exponential times. More precisely, let $`\mathrm{\Lambda }^d`$ be a smooth domain and set $`\mathrm{\Lambda }_N=N\mathrm{\Lambda }^d`$. We consider a Markov process on the state space $`X^{\mathrm{\Lambda }_N}`$, where $`X`$ is a subset of $``$, e.g. $`X=\{0,1\}`$ when an exclusion principle is imposed. The number of particles at site $`x\mathrm{\Lambda }_N`$ is denoted by $`\eta _xX`$ and the whole configuration by $`\eta X^{\mathrm{\Lambda }_N}`$. The dynamics is specified by a continuous time Markov process on the state space $`X^{\mathrm{\Lambda }_N}`$ with infinitesimal generator $`L_N=N^2\left[L_{0,N}+L_{b,N}\right]`$ defined as follows: for functions $`f:X^{\mathrm{\Lambda }_N}`$, $$\begin{array}{ccc}\hfill L_{0,N}f(\eta )& =& \frac{1}{2}\underset{\begin{array}{c}x,y\mathrm{\Lambda }_N\\ |xy|=1\end{array}}{}c_{x,y}(\eta )\left[f(\sigma ^{x,y}\eta )f(\eta )\right],\hfill \\ \hfill L_{b,N}f(\eta )& =& \frac{1}{2}\underset{\begin{array}{c}x\mathrm{\Lambda }_N,y\mathrm{\Lambda }_N\\ |xy|=1\end{array}}{}\left\{c_{x,y}(\eta )\left[f(\sigma ^{x,y}\eta )f(\eta )\right]+c_{y,x}(\eta )\left[f(\sigma ^{y,x}\eta )f(\eta )\right]\right\}.\hfill \end{array}$$ (2.1) Here $`|x|`$ stands for the usual Euclidean norm and $`\sigma ^{x,y}\eta `$, $`x,y\mathrm{\Lambda }_N`$, for the configuration obtained from $`\eta `$ by moving a particle from $`x`$ to $`y`$: $$\left(\sigma ^{x,y}\eta \right)_z=\{\begin{array}{ccc}\eta _z\hfill & \text{ if }& zx,y\hfill \\ \eta _y+1\hfill & \text{ if }& z=y\hfill \\ \eta _x1\hfill & \text{ if }& z=x.\hfill \end{array}$$ If $`x\mathrm{\Lambda }_N`$, $`y\mathrm{\Lambda }_N`$, then $`\sigma ^{y,x}\eta `$ is obtained from $`\eta `$ by creating a particle at $`x`$, while $`\sigma ^{x,y}\eta `$ is obtained by annihilating a particle at $`x`$. Therefore the generator $`L_{0,N}`$ describes the bulk dynamics which preserves the total number of particles whereas $`L_{b,N}`$ models the particle reservoirs at the boundary of $`\mathrm{\Lambda }_N`$. Note that we already speeded up the microscopic time by $`N^2`$ in the definition of $`L_N`$, which corresponds to the diffusive scaling. Assume that the bulk rates $`c_{x,y}`$, $`x,y\mathrm{\Lambda }_N`$, satisfy the local detailed balance with respect to a Gibbs measure defined by a Hamiltonian $``$ and in presence of an external vector field $`E=(E_1,\mathrm{},E_d)`$ smooth on the macroscopic scale. Likewise, assume that the boundary rates $`c_{x,y}`$, $`c_{y,x}`$, $`x\mathrm{\Lambda }_N`$, $`y\mathrm{\Lambda }_N`$, satisfy the local detailed balance with respect to $``$ and in presence of a chemical potential $`\lambda _0(y/N)`$ smooth on the macroscopic scale. The above requirements are met by the following formal definitions. Fix a smooth function $`\lambda _0:\mathrm{\Lambda }`$ and a Hamiltonian $``$. Consider jump rates $`c_{x,y}^0`$ satisfying the detailed balance with respect to the Gibbs measure associated to $``$ with free boundary conditions if $`x,y\mathrm{\Lambda }_N`$, while if $`x\mathrm{\Lambda }_N`$, $`y\mathrm{\Lambda }_N`$ we add the chemical potential $`\lambda _0(y/N)`$: $$\begin{array}{c}c_{x,y}^0(\eta )=\mathrm{exp}\left\{(\sigma ^{x,y}\eta )(\eta )\right\}c_{y,x}^0(\sigma ^{x,y}\eta ),x,y\mathrm{\Lambda }_N;\hfill \\ c_{x,y}^0(\eta )=\mathrm{exp}\left\{(\sigma ^{x,y}\eta )(\eta )+\lambda _0(y/N)\right\}c_{y,x}^0(\sigma ^{x,y}\eta ),x\mathrm{\Lambda }_N,y\mathrm{\Lambda }_N.\hfill \end{array}$$ Note that we included the inverse temperature in the Hamiltonian $``$. Of course if $`\eta _x=0`$ then $`c_{x,y}^0(\eta )=0`$. Fix a smooth vector field $`E=(E_1,\mathrm{},E_d):\mathrm{\Lambda }^d`$ and let $$c_{x,x+e_i}(\eta ):=e^{N^1E_i(x/N)}c_{x,x+e_i}^0(\eta ),c_{x+e_i,x}(\eta ):=e^{N^1E_i(x/N)}c_{x+e_i,x}^0(\eta ),$$ (2.2) where $`\{e_1,\mathrm{},e_d\}`$ stands for the canonical basis in $`^d`$. Namely, for $`N`$ large, by expanding the exponential, particles at site $`x`$ feel a drift $`N^1E(x/N)`$. Typically, for a non equilibrium model, we would consider $`\mathrm{\Lambda }`$ as the $`d`$-dimensional cube of side one, the system under a constant field $`E/N`$ and a chemical potential $`\lambda _0`$ satisfying $`\lambda _0(y/N)=\gamma _0`$ if the first coordinate of $`y`$ is $`0`$, $`\lambda _0(y/N)=\gamma _1`$ if the first coordinate of $`y`$ is $`N`$, imposing periodic boundary conditions in the other directions of $`\mathrm{\Lambda }`$. By setting $`c_{x,y}=0`$ if both $`x`$ and $`y`$ do not belong to $`\mathrm{\Lambda }_N`$, we can rewrite the full generator $`L_N`$ as follows $$L_Nf(\eta )=\frac{N^2}{2}\underset{\begin{array}{c}x,y^d\\ |xy|=1\end{array}}{}c_{x,y}(\eta )\left[f(\sigma ^{x,y}\eta )f(\eta )\right]$$ (2.3) We consider an initial condition $`\eta X^{\mathrm{\Lambda }_N}`$. The trajectory of the Markov process $`\eta (t)`$, $`t0`$, is an element on the path space $`D(_+;X^{\mathrm{\Lambda }_N})`$, which consists of piecewise constant paths with values in $`X^{\mathrm{\Lambda }_N}`$. We shall denote by $`_\eta ^N`$ the probability measure on $`D(_+;X^{\mathrm{\Lambda }_N})`$ corresponding to the distribution of the process $`\eta (t)`$, $`t0`$ with initial condition $`\eta `$. Examples of stochastic lattices gases are the simple exclusion processes in which $`X=\{0,1\}`$, $`=0`$ and $`c_{x,y}^0(\eta )=\eta _x[1\eta _y]`$ and zero range processes in which $`X=`$, $`(\eta )=_x_{1k\eta (x)}\mathrm{log}g(k)`$, for some function $`g:_+`$ such that $`g(0)=0`$, and $`c_{x,y}^0(\eta )=g(\eta _x)`$. 3. Macroscopic description of lattice gases The empirical density $`\pi ^N`$ can be naturally defined as follows. To each microscopic configuration $`\eta X^{\mathrm{\Lambda }_N}`$ we associate a macroscopic profile $`\pi ^N(u)`$, $`u\mathrm{\Lambda }`$, by requiring that for any smooth function $`G:\mathrm{\Lambda }`$ $$\pi ^N,G=_\mathrm{\Lambda }𝑑u\pi ^N(u)G(u)=\frac{1}{N^d}\underset{x\mathrm{\Lambda }_N}{}G(x/N)\eta _x$$ (3.1) so that $`\pi ^N(u)`$ is the local density at the macroscopic point $`u=x/N`$ in $`\mathrm{\Lambda }`$. Of course $`\pi ^N(u)`$ is really a sum of point masses at the points $`x/N`$ with weight $`\eta _x/N^d`$; in the limit $`N\mathrm{}`$ it will however weakly converge to a “true” function $`\rho (u)`$. The definition of the empirical current is slightly more complicated. Indeed it is not a function of the configuration $`\eta X^{\mathrm{\Lambda }_N}`$ but of the trajectory $`\{\eta (t)\}_{t0}D(_+;X^{\mathrm{\Lambda }_N})`$. Given an oriented bond $`(x,y)`$, let $`𝒩^{x,y}(t)`$ be the number of particles that jumped from $`x`$ to $`y`$ in the time interval $`[0,t]`$. Here we adopt the convention that $`𝒩^{x,y}(t)`$ is the number of particles created at $`y`$ due to the reservoir at $`x`$ if $`x\mathrm{\Lambda }_N`$, $`y\mathrm{\Lambda }_N`$ and that $`𝒩^{x,y}(t)`$ represents the number of particles that left the system at $`x`$ by jumping to $`y`$ if $`x\mathrm{\Lambda }_N`$, $`y\mathrm{\Lambda }_N`$. The difference $`Q^{x,y}(t)=𝒩^{x,y}(t)𝒩^{y,x}(t)`$ is the net number of particles flown across the bond $`\{x,y\}`$ in the time interval $`[0,t]`$. Given a trajectory $`\eta (s)`$, $`0st`$, the instantaneous current across $`\{x,y\}`$ is defined as $`dQ_t^{x,y}/dt`$. This is a sum of $`\delta `$–functions localized at the jump times with weight $`+1`$, resp. $`1`$, if a particle jumped from $`x`$ to $`y`$, resp. from $`y`$ to $`x`$. For a given realization of the process $`\eta (t)`$ in $`D(_+;X^{\mathrm{\Lambda }_N})`$, we define the corresponding empirical current $`𝒥^N`$ as follows. Let $`T>0`$ and pick a smooth vector field $`G=(G_1,\mathrm{},G_d)`$ defined on $`[0,T]\times \mathrm{\Lambda }`$. We then set $`𝒥^N,G_T`$ $`=`$ $`{\displaystyle _0^T}𝑑t{\displaystyle _\mathrm{\Lambda }}𝑑uG(t,u)𝒥^N(t,u)`$ (3.2) $`=`$ $`{\displaystyle \frac{1}{N^{d+1}}}{\displaystyle \underset{i=1}{\overset{d}{}}}{\displaystyle \underset{x}{}}{\displaystyle _0^T}G_i(t,x/N)𝑑Q^{x,x+e_i}(t),`$ where $``$ stands for the inner product in $`^d`$ and we sum over all $`x`$ such that either $`x\mathrm{\Lambda }_N`$ or $`x+e_i\mathrm{\Lambda }_N`$. The empirical current $`𝒥^N`$ is therefore a signed measure on $`\left([0,T]\times \mathrm{\Lambda }\right)^d`$, while we recall that the empirical density is a positive measure on $`\mathrm{\Lambda }`$. The normalization $`N^{(d+1)}`$ in (3.2) has been chosen so that the empirical current has a finite limit as $`N\mathrm{}`$. The local conservation of the number of particles is expressed by $$\eta _x(t)\eta _x(0)+\underset{y:|xy|=1}{}Q^{x,y}(t)=0.$$ It gives the following continuity equation for the empirical density and current. Let $`G`$ be a smooth function on a neighborhood of the closure of $`\mathrm{\Lambda }`$. Denote by $`_NG`$ the vector field whose coordinates are $`\left(_NG\right)_i(u)=N[G(u+e_i/N)G(u)]`$. Then $$\pi ^N(T),G\pi ^N(0),G=𝒥^N,_NG_T_0^T\frac{1}{N^d}\underset{\begin{array}{c}x\mathrm{\Lambda }_N,y\mathrm{\Lambda }_N\\ |xy|=1\end{array}}{}G(y/N)dQ^{x,y}(t)$$ The above equation can be formally stated as the continuity equation $$_t\pi ^N+_N𝒥^N=0$$ (3.3) In particular, given the initial condition, the trajectory of the process, described by the empirical density $`\pi ^N`$ can be completely recovered from the empirical current $`𝒥^N`$. We briefly discuss at the heuristic level the law of large numbers, as $`N\mathrm{}`$, for the empirical density and the empirical current. Details are given in Appendix 6.4. Fix a sequence of configurations $`\eta ^N`$ and assume that its associated empirical measure $`\pi ^N`$ converges to $`\rho _0(u)du`$ for some density profile $`\rho _0:\mathrm{\Lambda }_+`$. Let us denote by $`\rho =\rho (t,u)`$, $`J=J(t,u)`$, the limiting values of $`\pi ^N(t,u)`$, $`𝒥^N(t,u)`$, respectively. Here $`\pi ^N(t,u)`$ is the empirical density associated to the configuration $`\eta (t)`$ and $`𝒥^N(t,u)`$ has been defined in (3.2). The microscopic relation (3.3) implies the continuity equation $$_t\rho +J=0$$ (3.4) To derive a closed evolution for $`\rho `$ and $`J`$, we need to express the current $`J`$ in terms of the density $`\rho `$. To simplify the exposition, we assume the process to be gradient: there exist local functions $`h_0^{(i)}(\eta )`$, $`i=1,\mathrm{},d`$, depending on the configuration $`\eta `$ around $`0`$, so that for any $`i=1,\mathrm{},d`$ $$c_{x,x+e_i}^0(\eta )c_{x+e_i,x}^0(\eta )=h_x^{(i)}(\eta )h_{x+e_i}^{(i)}(\eta )$$ where $`h_x^{(i)}`$ is the function $`h_0^{(i)}`$ evaluated on the configuration $`\eta `$ translated by $`x`$. Denote by $`\mu _\lambda `$ the infinite volume grand canonical ensemble relative to the Hamiltonian $``$ with chemical potential $`\lambda `$. Choose the chemical potential $`\lambda =\lambda (\rho )`$ so that $`\mu _\lambda [\eta _0]=\rho `$ and define $$d^{(i)}(\rho )=\mu _{\lambda (\rho )}\left[h_0^{(i)}\right],\chi ^{(i)}(\rho ):=(1/2)\mu _{\lambda (\rho )}\left[c_{0,e_i}^0+c_{e_i,0}^0\right]$$ (3.5) We show in Appendix 6.4 that the current $`J`$ can be expressed in terms of the density $`\rho `$ as $$J=\frac{1}{2}D(\rho )\rho +\chi (\rho )E=:J(\rho )$$ (3.6) where $`D`$ and $`\chi `$ are $`d\times d`$ diagonal matrices with entries $`D_{ii}(\rho )=\frac{d}{d\rho }d^{(i)}(\rho )`$ and $`\chi _{ii}(\rho )=\chi ^{(i)}(\rho )`$. For non gradient systems the diffusion matrix $`D`$ and the mobility $`\chi `$ are not in general diagonal. In such a situation $`D`$ is given by a Green–Kubo formula \[31, II.2.2\] and $`\chi `$ can be obtained by linear response theory \[31, II.2.5\]. These coefficients are related by Einstein relation $`D=R^1\chi `$, where $`R`$ is the compressibility: $`R^1=F_0^{\prime \prime }`$, in which $`F_0`$ is the equilibrium free energy associated to the Hamiltonian $``$, . To conclude the description of the evolution, it remains to examine the evolution at the boundary of $`\mathrm{\Lambda }`$. We claim that the density is fixed there because we speeded up diffusively the non-conservative Glauber dynamics at the boundary: $$\lambda \left(\rho (t,u)\right)=\lambda _0(u)u\mathrm{\Lambda }$$ (3.7) The macroscopic evolution of the density and the current is thus described by the equation $$\{\begin{array}{c}_t\rho +J=0,u\mathrm{\Lambda }\hfill \\ J=\frac{1}{2}D(\rho )\rho +\chi (\rho )E,u\mathrm{\Lambda }\hfill \\ \lambda \left(\rho (t,u)\right)=\lambda _0(u),u\mathrm{\Lambda }\hfill \\ \rho (0,)=\rho _0().\hfill \end{array}$$ The stationary density profile $`\overline{\rho }=\overline{\rho }(u)`$, $`u\mathrm{\Lambda }`$, is the stationary solution of the hydrodynamic equation, that is $$\{\begin{array}{c}J(\overline{\rho }(u))=0,u\mathrm{\Lambda }\hfill \\ \lambda \left(\overline{\rho }(u)\right)=\lambda _0(u)u\mathrm{\Lambda },\hfill \end{array}$$ If we let the macroscopic time diverge, $`t\mathrm{}`$, $`\rho (t)\overline{\rho }`$ and $`J(\rho (t))`$ converges to $`J(\overline{\rho })`$, which is the current maintained by the stationary state. We next discuss the large deviation properties of the empirical current. More details are given in Appendix 6.4. As before we consider a sequence of initial configuration $`\eta ^N`$ such that the empirical density $`\pi ^N(\eta ^N)`$ converges to some density profile $`\rho _0`$. We fix a smooth vector field $`j:[0,T]\times \mathrm{\Lambda }^d`$. The large deviation principle for the current states that $$_{\eta ^N}^N\left(𝒥^N(t,u)j(t,u),(t,u)[0,T]\times \mathrm{\Lambda }\right)\mathrm{exp}\left\{N^d_{[0,T]}(j)\right\}$$ (3.8) where the rate functional $``$ is $$_{[0,T]}(j)=\frac{1}{2}_0^T𝑑t[j(t)J(\rho (t))],\chi (\rho (t))^1[j(t)J(\rho (t))]$$ (3.9) in which $`\rho (t)=\rho (t,u)`$ is obtained by solving the continuity equation $$\{\begin{array}{c}_t\rho (t,u)+j(t,u)=0\hfill \\ \rho (0,u)=\rho _0(u)\hfill \end{array}$$ (3.10) and $`J(\rho )`$ is given by (3.6). Of course there are compatibility conditions to be satisfied, for instance if we have chosen a $`j`$ such that $`\rho (t,u)`$ becomes negative for some $`(t,u)[0,T]\times \mathrm{\Lambda }`$ then $`_{[0,T]}(j)=+\mathrm{}`$. Notice that, even if not indicated explicitly in the notation, the rate functional $``$ depends on the initial density profile $`\rho _0`$, through equation (3.10). We note that in the large deviation functional (3.9) the fluctuation of the density $`\rho (t)`$ is determined by the current $`j(t)`$. The large deviations properties of the density, which we described in for non equilibrium stochastic lattice gases, can thus be deduced from the ones of the current, see Appendix 6.4 for the details. This is due to the fact that the continuity equation, as already remarked, holds exactly at the microscopic level, see (3.3). On the other hand the constitutive equation (3.6) holds only in the limit $`N\mathrm{}`$ when fluctuations can be neglected. 4. Large deviation of the time averaged current We want to study the fluctuations of the time average of the empirical current over a large time interval $`[0,T]`$; the corresponding probability can be obtained from the space time large deviation principle (3.8). Fix $`T>0`$ and a divergence free vector field $`J=J(u)`$. We introduce the set of possible paths $`j`$ of the current with time average $`J`$ $$𝒜_{T,J}=\{j=j(t,u):\frac{1}{T}_0^T𝑑tj(t,u)=J(u)\}$$ The condition of vanishing divergence on $`J`$ is required by the local conservation of the number of particles. By the large deviations principle (3.8), for $`T`$ and $`N`$ large we have $`_{\eta ^N}^N\left({\displaystyle \frac{1}{T}}{\displaystyle _0^T}𝑑t𝒥^N(t)J\right)\mathrm{exp}\left\{N^dT\mathrm{\Phi }(J)\right\}`$ (4.1) where the logarithmic equivalence is understood by sending first $`N\mathrm{}`$ and then $`T\mathrm{}`$. In Subsection 6.4 below we shall show that for the zero range process the limits can be taken in the opposite order; we expect this to be true in general. The functional $`\mathrm{\Phi }`$ is given by $$\mathrm{\Phi }(J)=\underset{T\mathrm{}}{lim}\underset{j𝒜_{T,J}}{inf}\frac{1}{T}_{[0,T]}(j)=\underset{T>0}{inf}\underset{j𝒜_{T,J}}{inf}\frac{1}{T}_{[0,T]}(j)$$ (4.2) By a standard sub–additivity argument we show that the limit $`T\mathrm{}`$ exists and coincides with the infimum in $`T`$. Indeed, given $`j_1𝒜_{T,J}`$ and $`j_2𝒜_{S,J}`$, we have $$_{[0,T+S]}(j)=_{[0,T]}(j_1)+_{[0,S]}(j_2)$$ (4.3) where $`j`$ is obtained by gluing $`j_1`$ and $`j_2`$. Here we used the invariance of $``$ under time shift and that $`j_1𝒜_{T,J}`$ implies that the corresponding density $`\rho _1`$, obtained by solving the continuity equation (3.10), satisfies $`\rho _1(0)=\rho _1(T)`$. From the previous equation we get the sub–additivity property: $$\underset{j𝒜_{T+S,J}}{inf}_{[0,T+S]}(j)\underset{j𝒜_{T,J}}{inf}_{[0,T]}(j)+\underset{j𝒜_{S,J}}{inf}_{[0,S]}(j)$$ Even if the rate functional $``$ depends on the initial density profile $`\rho _0`$, by taking the limit in (4.2) it is easy to show $`\mathrm{\Phi }`$ does not. We now prove that $`\mathrm{\Phi }`$ is a convex functional. Let $`0<p<1`$ and $`J=pJ_1+(1p)J_2`$, we want to show that $`\mathrm{\Phi }(J)p\mathrm{\Phi }(J_1)+(1p)\mathrm{\Phi }(J_2)`$. By (4.2), given $`\epsilon >0`$ we can find $`T>0`$, $`j_1𝒜_{pT,J_1}`$, and $`j_2𝒜_{(1p)T,J_2}`$ so that $$\begin{array}{ccc}\mathrm{\Phi }(J_1)\hfill & & \frac{1}{pT}_{[0,pT]}(j_1)\epsilon \hfill \\ \mathrm{\Phi }(J_2)\hfill & & \frac{1}{(1p)T}_{[0,(1p)T]}(j_2)\epsilon \hfill \end{array}$$ By the same arguments used in (4.3), the path obtained by gluing $`j_1`$ with $`j_2`$, denoted by $`j`$, is in the set $`𝒜_{T,J}`$. Therefore, $$\mathrm{\Phi }(J)\frac{1}{T}_{[0,T]}(j)p\mathrm{\Phi }(J_1)+(1p)\mathrm{\Phi }(J_2)+\epsilon $$ which proves the convexity of $`\mathrm{\Phi }`$. These arguments are standard in proving the existence and convexity of thermodynamic functions in statistical mechanics. We next study the variational problem on the right hand side of (4.2). We begin by deriving an upper bound. Given $`\rho =\rho (u)`$ and $`J=J(u)`$, $`J=0`$, let us introduce the functionals $`𝒰(\rho ,J)`$ $`=`$ $`{\displaystyle \frac{1}{2}}JJ(\rho ),\chi (\rho )^1[JJ(\rho )]`$ (4.4) $`U(J)`$ $`=`$ $`\underset{\rho }{inf}𝒰(\rho ,J)`$ (4.5) where the minimum in (4.5) is carried over all profiles $`\rho `$ satisfying the boundary condition (3.7) and $`J(\rho )`$ is given by (3.6). When $`J`$ is constant, that is, in the one–dimensional case, the functional $`U`$ is the one introduced in . We claim that $$\mathrm{\Phi }(J)U(J).$$ (4.6) The strategy to prove this bound is quite simple, see also . Let $`\widehat{\rho }=\widehat{\rho }(J)`$ be the density profile which minimizes the variational problem (4.5). Given the initial density profile $`\rho _0`$, we choose some fixed time $`\tau >0`$ and a current $`\widehat{ȷ}=\widehat{ȷ}(u)`$ which moves the density from $`\rho _0`$ to $`\widehat{\rho }`$ in a time lag $`\tau `$, namely such that $`\tau \widehat{ȷ}=\rho _0\widehat{\rho }`$. We now construct the path $`j=j(t,u)`$, $`(t,u)[0,T]\times \mathrm{\Lambda }`$ as follows $$j(t)=\{\begin{array}{ccc}\widehat{ȷ}& \text{if}& 0t<\tau \hfill \\ \frac{T}{T2\tau }J& \text{if}& \tau t<T\tau \hfill \\ \widehat{ȷ}& \text{if}& T\tau tT\hfill \end{array}$$ The corresponding density $`\rho (t)`$ is obtained by solving the continuity equation (3.10), i.e. $$\rho (t)=\{\begin{array}{ccc}\rho _0+\frac{t}{\tau }(\widehat{\rho }\rho _0)& \text{if}& 0t<\tau \hfill \\ \widehat{\rho }& \text{if}& \tau t<T\tau \hfill \\ \rho _0+\frac{Tt}{\tau }(\widehat{\rho }\rho _0)& \text{if}& T\tau tT\hfill \end{array}$$ It is straightforward to verify that $`j𝒜_{T,J}`$, as well as $`lim_T\mathrm{}\frac{1}{T}_{[0,T]}(j)=U(J)`$. By the convexity of $`\mathrm{\Phi }(J)`$ we can improve the upper bound (4.6) for free. Let us denote by $`U^{}`$ the convex envelope of $`U`$, i.e. the largest convex functional below $`U`$. By taking the convex envelope in (4.6) we get $$\mathrm{\Phi }(J)U^{}(J)$$ (4.7) We next discuss a lower bound for the variational problem (4.2). We denote by $`\stackrel{~}{𝒰}`$ and $`\stackrel{~}{U}`$ the same functionals as in (4.4)–(4.5), but now defined on the space of all currents without the conditions of vanishing divergence. Let also $`\stackrel{~}{U}^{}`$ be the convex envelope of $`\stackrel{~}{U}`$. Let $`j𝒜_{T,J}`$. By the convexity of $`\stackrel{~}{U}^{}`$ in the set of all currents, we get $$\begin{array}{ccc}\frac{1}{T}_{[0,T]}(j)\hfill & =& \frac{1}{T}_0^T𝑑t\stackrel{~}{𝒰}(\rho (t),j(t))\frac{1}{T}_0^T𝑑t\stackrel{~}{U}(j(t))\hfill \\ & & \frac{1}{T}_0^T𝑑t\stackrel{~}{U}^{}(j(t))\stackrel{~}{U}^{}(J)\hfill \end{array}$$ which implies $$\mathrm{\Phi }(J)\stackrel{~}{U}^{}(J)$$ (4.8) The upper and lower bounds (4.7) and (4.8) are different in general. For a divergence free $`J`$ we have $`\stackrel{~}{U}(J)=U(J)`$ but since the convex envelopes are considered in different spaces, we only have $`\stackrel{~}{U}^{}(J)U^{}(J)`$. The derivation of the upper bound shows that our result differs from the one in if $`U`$ is not convex. Moreover, if $`\mathrm{\Phi }(J)<U(J)`$, the optimal density path $`\rho `$ in the variational problem (4.2) must be time dependent. We now examine how different behaviors of the solution to the variational problem (4.2) reflect different dynamical regimes that we interpret as dynamical phase transitions. It is convenient to work in the time interval $`[T,T]`$ instead of $`[0,T]`$. We consider the system in the ensemble defined by conditioning on the event $`(2T)^1_T^T𝑑t𝒥^N(t)=J`$ with $`N`$ and $`T`$ large. The parameter $`J`$ plays therefore the role of an intensive thermodynamic variable and the convexity of $`\mathrm{\Phi }`$ expresses a stability property with respect to variations of $`J`$. If $`\mathrm{\Phi }(J)=U(J)`$ and the minimum for (4.5) is attained for $`\rho =\widehat{\rho }(J)`$ we have a state analogous to a unique phase: by observing the system at any fixed time $`t`$ we see, with probability converging to one as $`N,T\mathrm{}`$, the density $`\pi ^N(t)\widehat{\rho }(J)`$ and the current $`𝒥^N(t)J`$. When $`\mathrm{\Phi }(J)=U^{}(J)<U(J)`$, we have a state analogous to a phase coexistence. Suppose for example $`J=pJ_1+(1p)J_2`$ and $`U(J)>U^{}(J)=pU(J_1)+(1p)U(J_2)`$ for some $`p,J_1,J_2`$. The values $`p,J_1,J_2`$ are determined by $`J`$ and $`U`$. The density profile is then not determined, but rather we observe with probability $`p`$ the profile $`\widehat{\rho }(J_1)`$ and with probability $`1p`$ the profile $`\widehat{\rho }(J_2)`$. Actually there is a memory of initial condition: if we take $`\rho (T)=\rho (T)=\widehat{\rho }(J_1)`$ we will see the density $`\widehat{\rho }(J_1)`$ in the time intervals $`[T,(1p)T]`$ and $`[(1p)T,T]`$, $`\widehat{\rho }(J_2)`$ in $`[(1p)T,(1p)T]`$. Consider now the case in which a minimizer for (4.2) is a function $`\widehat{ȷ}(t)`$ not constant in $`t`$. This is possible (an example will be given in Subsection 6.2) only when $`\mathrm{\Phi }(J)<U^{}(J)`$. Suppose first that $`\widehat{ȷ}(t)`$ is periodic with period $`\tau `$ and denote by $`\widehat{\rho }(t)`$ the corresponding density. Of course we have $`\tau ^1_0^\tau 𝑑t\widehat{ȷ}(t)=J`$. In such a case we have in fact a one parameter family of minimizers which are obtained by a time shift $`\alpha [0,\tau ]`$. By choosing $`2T`$ an integral multiple of $`\tau `$ and $`\rho (T)=\widehat{\rho }(\alpha )`$ for some $`\alpha [0,\tau ]`$ then the empirical density in the conditional ensemble will follow the path $`\widehat{\rho }(t+\alpha +T)`$. This behavior is analogous to a non translation invariant state in equilibrium statistical mechanics, like a crystal. Finally if $`\widehat{ȷ}(t)`$ is time dependent and not periodic the corresponding state is analogous to a quasi–crystal. The asymptotic (4.1) can be formulated in terms of the Laplace transform of the empirical current as follows. For each divergence free, time independent, vector field $`\lambda =\lambda (u)`$ we have $$\underset{T\mathrm{}}{lim}\underset{N\mathrm{}}{lim}\frac{1}{TN^d}\mathrm{log}𝔼_{\eta ^N}^N\left(e^{N^d𝒥^N,\lambda _T}\right)=\mathrm{\Phi }^{}(\lambda )$$ (4.9) where $`\mathrm{\Phi }^{}(\lambda )`$ is the Legendre transform of $`\mathrm{\Phi }(J)`$: $$\mathrm{\Phi }^{}(\lambda )=\underset{J}{sup}\left\{\lambda ,J\mathrm{\Phi }(J)\right\},$$ where the supremum is carried over all the divergence free vector fields $`J`$. It follows from (4.6) that $`U^{}\mathrm{\Phi }^{}`$. We conclude this section deriving a variational expression for $`U^{}`$. Recall the definitions (4.4), (4.5) of $`U`$. $$\begin{array}{ccc}\hfill U^{}(\lambda )& =& \underset{J,\rho }{sup}\left\{\lambda ,J\frac{1}{2}[JJ(\rho )],\chi (\rho )^1[JJ(\rho )]\right\}\hfill \\ & =& \underset{J,\rho }{sup}\{\frac{1}{2}[JJ(\rho )\chi (\rho )\lambda ],\chi (\rho )^1[JJ(\rho )\chi (\rho )\lambda ]\hfill \\ & & +\frac{1}{2}\lambda ,\chi (\rho )\lambda +\lambda ,J(\rho )\}\hfill \end{array}$$ To compute the supremum over $`J`$ we decompose the vector field $`J(\rho )+\chi (\rho )\lambda `$ as follows $$J(\rho )+\chi (\rho )\lambda =\chi (\rho )\psi +\left[J(\rho )+\chi (\rho )\left(\lambda \psi \right)\right]$$ (4.10) where $`\psi `$ solves $$\{\begin{array}{cc}\left(\chi (\rho )\psi \right)=(J(\rho )+\chi (\rho )\lambda )\hfill & u\mathrm{\Lambda }\hfill \\ \psi (u)=0\hfill & u\mathrm{\Lambda }\hfill \end{array}$$ Since the second term in the decomposition (4.10) is divergence free we get $$U^{}(\lambda )=\underset{\rho }{sup}\left\{\frac{1}{2}\psi ,\chi (\rho )\psi +\frac{1}{2}\lambda ,\chi (\rho )\lambda +\lambda ,J(\rho )\right\}$$ (4.11) where the supremum is over all density profiles $`\rho `$ satisfying $`F_0^{}(\rho (u))=\lambda _0(u)`$, $`u\mathrm{\Lambda }`$. 5. Time–reversal and Gallavotti–Cohen symmetry In this Section we discuss the properties of the rate functional for the current under time reversal. We also show that the functional $`\mathrm{\Phi }`$, which measures the probability of deviations of the time averaged current, satisfies a fluctuation theorem analogous to the Gallavotti–Cohen symmetry. ## 5.1 Time–reversal properties of the rate functional In the previous Sections we discussed a large deviation principle given a fixed initial condition $`\eta ^N`$ associated to a density profile $`\rho _0`$, i.e. $`\pi ^N(\eta ^N)\rho _0`$. Now we consider instead the stationary process, namely the initial condition is distributed according to the invariant measure $`\mu ^N`$ which is defined by $`_\eta \mu ^N(\eta )L_Nf(\eta )=0`$ for any observable $`f:X^{\mathrm{\Lambda }_N}`$; recall the generator $`L_N`$ has been defined in (2.3). As discussed in the Introduction, the large deviations of the empirical density under the distribution $`\mu ^N`$ are described by the non equilibrium free energy $`F`$, i.e., $$\mu ^N\left(\pi ^N\rho \right)\mathrm{exp}\left\{N^dF(\rho )\right\}$$ (5.1) In we show that the functional $`F`$, which for equilibrium states is trivially related to the free energy, can be characterized by a variational problem on the dynamical rate functional for the density $``$ introduced in (1.4), see also (A.15). To this variational problem is associated a Hamilton–Jacobi equation which plays a crucial role. In order to analyze the large deviations properties of the stationary process, since the initial condition is not fixed, it is natural to consider the joint fluctuations of the empirical density and current. We have $$_{\mu ^N}^N\left(\pi ^N\rho ,𝒥^Njt[T,T]\right)\mathrm{exp}\{N^d𝒢_{[T,T]}(\rho ,j)\}$$ (5.2) Here $`_{\mu ^N}^N`$, a probability measure on the space $`D(;X^{\mathrm{\Lambda }_N})`$, is the stationary process. Of course, fluctuations of the density and of the current are not independent since the continuity equation $`_t\rho +j=0`$ must be satisfied. Therefore the large deviation functional is $$𝒢_{[T,T]}(\rho ,j)=\{\begin{array}{cc}F(\rho (T))+_{[T,T]}(j)\hfill & \text{if }_t\rho +j=0\hfill \\ +\mathrm{}\hfill & \text{otherwise}\hfill \end{array}$$ (5.3) where $``$ has been introduced in (3.9). If we are interested only in the current fluctuations in the stationary process we get the appropriate rate functional by projecting (5.3), $$\underset{\rho _0}{inf}\left\{F(\rho _0)+_{[T,T]}(j)\right\}.$$ Let us denote by $`L_N^a`$ the adjoint of the generator $`L_N`$ (2.3) with respect to the invariant measure $`\mu ^N`$. We call the process generated by $`L_N^a`$, which is still Markovian, the adjoint process. We remark that the invariant measure of the adjoint process is again $`\mu ^N`$. Given a path $`\eta D(;X^{\mathrm{\Lambda }_N})`$ its time reversed is naturally defined as $`[\vartheta \eta ](t)=\eta (t)`$. The stationary adjoint process, that we denote by $`_{\mu ^N}^{N,a}`$, is the time reversal of $`_{\mu ^N}^N`$, i.e. we have $`_{\mu ^N}^{N,a}=_{\mu ^N}^N\vartheta ^1`$. We extend the definition of the time reversal operator $`\vartheta `$ to the current as $`[\vartheta j](t)=j(t)`$. Note that the current $`j`$ changes sign under time–reversal. Then $$_{\mu ^N}^N\left(\pi ^N\rho ,𝒥^Njt[T,T]\right)=_{\mu ^N}^{N,a}\left(\pi ^N\vartheta \rho ,𝒥^N\vartheta jt[T,T]\right)$$ At the level of large deviations this implies $$𝒢_{[T,T]}(\rho ,j)=𝒢_{[T,T]}^a(\vartheta \rho ,\vartheta j)$$ (5.4) where $`𝒢_{[T,T]}^a`$ is the large deviation functional for the adjoint process. The relationship (5.4) has far reaching consequences. We next show that it implies a fluctuation dissipation relation for the current. We assume that the adjoint process has a dynamical large deviations principle of the same form as (5.3) with $``$ replaced by $`^a`$ where $$_{[T,T]}^a(j)=\frac{1}{2}_T^T𝑑t[j(t)J^a(\rho (t))],\chi (\rho (t))^1[j(t)J^a(\rho (t))],$$ in which $`J^a(\rho )`$ is the typical value of the current of the adjoint process. We divide both sides of (5.4) by $`2T`$ and take the limit $`T0`$. By using (5.3) we get $$\frac{\delta F}{\delta \rho },_t\rho =\frac{1}{2}jJ(\rho ),\chi (\rho )^1[jJ(\rho )]\frac{1}{2}j+J^a(\rho ),\chi (\rho )^1[j+J^a(\rho )]$$ recalling that $`_t\rho +j=0`$, this is equivalent to $$\frac{\delta F}{\delta \rho },j=J(\rho )+J^a(\rho ),\chi (\rho )^1j+\frac{1}{2}J(\rho )+J^a(\rho ),\chi (\rho )^1[J(\rho )J^a(\rho )]$$ which has to be satisfied for any $`\rho `$ and $`j`$. By using that $`\delta F/\delta \rho `$ vanishes at the boundary of $`\mathrm{\Lambda }`$, see , we can integrate by parts the left hand side above and get $`J(\rho )+J^a(\rho )=\chi (\rho ){\displaystyle \frac{\delta F}{\delta \rho }}`$ (5.5) $`J(\rho ),\chi (\rho )^1J(\rho )=J^a(\rho ),\chi (\rho )^1J^a(\rho )`$ (5.6) Equation (5.5) is a fluctuation dissipation for the current analogous to the one for the density discussed in . It also extends the relationships between currents and thermodynamic forces, see e.g. , to a non equilibrium setting. By plugging (5.5) into (5.6) we also get another derivation of the Hamilton–Jacobi equation mentioned before, i.e. $$\frac{1}{2}\frac{\delta F}{\delta \rho },\chi (\rho )\frac{\delta F}{\delta \rho }+\frac{\delta F}{\delta \rho },J(\rho )=0$$ Let us now consider the variational problem (4.2) as well as the same problem for the functional $`^a`$, we denote by $`\mathrm{\Phi }^a`$ the corresponding functional. From (5.4) we get $$\mathrm{\Phi }(J)=\mathrm{\Phi }^a(J)$$ (5.7) For reversible process this symmetry states that the functional $`\mathrm{\Phi }`$ is even. Let us consider a path $`j(t)`$, $`t[T,T]`$ such that $`(2T)^1_T^T𝑑tj(t)=J`$ for some divergence free vector field $`J`$. Recalling (3.6) and that $`D(\rho )\chi (\rho )^1=F_0^{\prime \prime }(\rho )`$ we have $$\chi (\rho )^1J(\rho )=\frac{1}{2}F_0^{}(\rho )+E$$ Since $`F_0^{}(\rho (u))=\lambda _0(u)`$, $`u\mathrm{\Lambda }`$, by developing the square in (3.9) and integrating by parts we get $$\frac{1}{2T}𝒢_{[T,T]}(\rho ,j)=\frac{1}{2T}𝒢_{[T,T]}(\vartheta \rho ,\vartheta j)2J,E+_\mathrm{\Lambda }𝑑\mathrm{\Sigma }\lambda _0J\widehat{n}$$ (5.8) where $`d\mathrm{\Sigma }`$ is the surface measure on $`\mathrm{\Lambda }`$ and $`\widehat{n}`$ is the outward normal to $`\mathrm{\Lambda }`$. In particular this relation implies that if $`\widehat{\rho },\widehat{ȷ}`$ is an optimal path for the variational problem defining $`\mathrm{\Phi }(J)`$ then $`\vartheta \widehat{\rho },\vartheta \widehat{ȷ}`$ is an optimal path for the variational problem defining $`\mathrm{\Phi }(J)`$. By taking the limit $`T\mathrm{}`$ in (5.8) we get $$\mathrm{\Phi }(J)\mathrm{\Phi }(J)=\mathrm{\Phi }(J)\mathrm{\Phi }^a(J)=2J,E+_\mathrm{\Lambda }𝑑\mathrm{\Sigma }\lambda _0J\widehat{n}$$ (5.9) which is a Gallavotti–Cohen type symmetry in our space time dependent setup for macroscopic observables. Note that the right hand side of (5.9) is the power produced by the external field and the boundary reservoirs (recall $`E`$ is the external field and $`\lambda _0`$ the chemical potential of the boundary reservoirs). ## 5.2 Entropy production Recall that we denote by $`_{\mu ^N}^N`$ the stationary state and by $`_{\mu ^N}^{N,a}`$ its time reversed, i.e. the stationary adjoint process. In the context of Markov processes the Gallavotti–Cohen observable is defined as $$W_N(T)=\frac{1}{2TN^d}\mathrm{log}\frac{d_\mu ^{N,a}}{d_\mu ^N}|_{[T,T]}$$ (5.10) where the subscript means that we consider both distributions in the time interval $`[T,T]`$. We introduced the factor $`2TN^d`$ in order to discuss the asymptotic $`N,T\mathrm{}`$. As discussed in \[25, §2.4\], $`W_N(T)`$ can be interpreted as the microscopic production of the Gibbs entropy. For $`N`$ fixed and $`T\mathrm{}`$ the functional $`W_N`$ satisfies a large deviation principle with rate function $`f_N`$ namely, $$_{\mu ^N}^N\left(W_N(T)q\right)\mathrm{exp}\left\{2TN^df_N(q)\right\}$$ (5.11) In it is shown that $`W_N`$ satisfies the Gallavotti–Cohen symmetry, which states that the odd part of $`f_N`$ is linear with a universal coefficient: $`f_N(q)f_N(q)=q`$. An elementary computation, analogous to the one in , shows that, for the stochastic lattice gases as introduced in Section 1.3, we can express the functional $`W_N`$ in terms of the empirical current. More precisely, we have $$\begin{array}{c}W_N(T)=\frac{1}{2TN^d}\{\mathrm{log}\frac{\mu ^N(\eta (T))}{\mu ^N(\eta (T))}+(\eta (T))(\eta (T))\hfill \\ \frac{2}{N}\underset{j=1}{\overset{d}{}}\underset{x}{}E_j(x/N)Q^{x,x+e_j}([T,T])+\underset{\begin{array}{c}x\mathrm{\Lambda }_N\\ y\mathrm{\Lambda }_N\end{array}}{}\lambda _0(y/N)Q^{x,y}([T,T])\},\hfill \end{array}$$ where the summation is carried over all $`x`$ such that either $`x\mathrm{\Lambda }_N`$ or $`x+e_j\mathrm{\Lambda }_N`$. The previous equation can be understood as an entropy balance. Indeed, in the right hand side the first term is, for $`N`$ large, the difference of the non equilibrium free energy at times $`T`$ and $`T`$, the second is the difference of the energy and the third is the work done by the external field $`E`$ and the boundary reservoirs. Therefore, $`W_N`$ can be interpreted as the total entropy produced by the system in the time interval $`[T,T]`$. Recalling the definition of the empirical current $`𝒥^N`$, we can rewrite the above equation as $$\begin{array}{ccc}W_N(T)\hfill & =& \frac{1}{2T}\{\frac{1}{N^d}[\mathrm{log}\frac{\mu ^N(\eta (T))}{\mu ^N(\eta (T))}+(\eta (T))(\eta (T))]\hfill \\ & & +2𝒥^N,E_{[T,T]}\frac{1}{N^d}\underset{\begin{array}{c}x\mathrm{\Lambda }_N\\ y\mathrm{\Lambda }_N\end{array}}{}\lambda _0(y/N)Q^{x,y}([T,T])\}\hfill \end{array}$$ (5.12) We emphasize that, while the empirical current is a vector in $`^d`$, $`W_N(t)`$ is a scalar. From the previous expression it follows that, for any $`\delta >0`$ $$\underset{T\mathrm{}}{lim}\underset{N\mathrm{}}{lim}_{\eta ^N}^N\left(\left|W_N(T)2E,J(\overline{\rho })_\mathrm{\Lambda }𝑑\mathrm{\Sigma }\lambda _0J(\overline{\rho })\widehat{n}\right|>\delta \right)=0$$ (5.13) where we recall that $`J(\overline{\rho })=(1/2)D(\overline{\rho })\overline{\rho }+\chi (\overline{\rho })E`$ is the typical current. We note that as $`T\mathrm{}`$ we can neglected the first line on the r.h.s. of (5.12) because it is a boundary term. We thus define $$\stackrel{~}{W}_N(T)=\frac{1}{2T}\left\{2𝒥^N,E_{[T,T]}\frac{1}{N^d}\underset{\begin{array}{c}x\mathrm{\Lambda }_N\\ y\mathrm{\Lambda }_N\end{array}}{}\lambda _0(y/N)Q^{x,y}[T,T]\right\}$$ (5.14) which satisfies, as $`T\mathrm{}`$ with $`N`$ fixed, the large deviation estimate (5.11) with the same rate function $`f_N`$. On the other hand, since $`\stackrel{~}{W}_N(T)`$ is a function of the empirical current, we can apply the large deviation principle (5.2). We then get, by taking *first* the limit $`N\mathrm{}`$ and *then* $`T\mathrm{}`$, $$_{\mu ^N}^N\left(\stackrel{~}{W}_N(T)q\right)\mathrm{exp}\left\{2TN^df(q)\right\}$$ (5.15) where the rate function $`f`$ can be expressed in terms of the functional $``$ namely $$f(q)=\underset{T\mathrm{}}{lim}\underset{j_{T,q}}{inf}\frac{1}{2T}_{[T,T]}(j)$$ in which we introduced the set of currents $$_{T,q}:=\{j:\frac{1}{2T}\left[2_T^T𝑑tj(t),E_T^T𝑑t_\mathrm{\Lambda }𝑑\mathrm{\Sigma }\lambda _0j(t)\widehat{n}\right]=q\}$$ (5.16) where we recall $`d\mathrm{\Sigma }`$ is the surface measure on $`\mathrm{\Lambda }`$ and $`\widehat{n}`$ is the outward normal to $`\mathrm{\Lambda }`$. Finally, since $`E`$ and $`\lambda _0`$ are time independent we can take the time average of the empirical current in (5.16). Recalling (4.2), we get $$f(q)=\underset{J_q}{inf}\mathrm{\Phi }(J)$$ (5.17) where $$_q:=\{J:\mathrm{\hspace{0.25em}2}J,E_\mathrm{\Lambda }𝑑\mathrm{\Sigma }\lambda _0J\widehat{n}=q,J=0\}$$ (5.18) where we inserted the condition $`J=0`$ because other things do not happen. The content of the variational problem (5.17) is to look for, among all possible currents, the best one to have a fixed entropy production. It is straightforward to verify that the symmetry (5.9) implies the classical Gallavotti–Cohen symmetry for the limiting functional $`f`$, i.e. $`f(q)f(q)=q`$. On the other hand, if $`d>1`$, equation (5.9) is more general than the classical Gallavotti–Cohen symmetry. In the above argument we first took the limit $`N\mathrm{}`$ and next $`T\mathrm{}`$, but we expect that these limits could be taken in any order. In particular these would imply $`lim_N\mathrm{}f_N(q)=f(q)`$. In Section 6.4 we prove that this is the case for the zero range process. We finally note that in the one–dimensional case, setting $`\mathrm{\Lambda }=[0,1]`$, we can easily solve (5.17). We get $$f(q)=\mathrm{\Phi }\left(\frac{q}{2E[\lambda _0(1)\lambda _0(0)]}\right)$$ 6. Dynamical phase transitions: examples As we have discussed in Section 1.3, we always have the following inequalities $$\stackrel{~}{U}^{}(J)\mathrm{\Phi }(J)U^{}(J)U(J)$$ (6.1) for any divergence free $`J`$. A natural question is when the above inequalities are strict and when are equalities, in particular when $`\mathrm{\Phi }=U`$. As discussed in Section 1.3, the strict inequality $`\mathrm{\Phi }(J)<U(J)`$ is a dynamical phase transition on the ensemble defined by conditioning on the event in which the time average current equals $`J`$. In this Section we discuss several examples which show that different scenarios actually do take place in concrete models. As we have shown in Section 1.3 the macroscopic behavior (including the probability of large fluctuations) of the system is determined by the transport coefficients $`D(\rho )`$ and $`\chi (\rho )`$. In this Section we consider these as given functions and discuss the properties of the variational problem defining $`\mathrm{\Phi }`$. Specific choices of $`D`$ and $`\chi `$ correspond to well studied microscopic models, such as the simple exclsion processes, the zero range process, and the KMP model . In Subsection 6.1 we find sufficient conditions on $`D`$ and $`\chi `$ implying $`\mathrm{\Phi }=U`$. In Subsection 6.2 we discuss periodic boundary conditions: under appropriate conditions on the transport coefficient and the external field we show that the minimizer for the variational problem defining $`U`$ is obtained when $`\rho `$ is constant (in space). Moreover, by considering travelling waves, we find for $`J`$ large a better (space-time dependent) strategy so that $`\mathrm{\Phi }<U`$. These conditions hold in particular for the KMP model with no external field. Moreover, for the exclusion process with sufficiently large external field, we show that there exists a travelling wave path of current whose cost is strictly less than the constant (in time and space) one. This was first observed in . In Subsection 6.3, we give an example where $`U`$ is non convex which implies $`\mathrm{\Phi }<U`$. Finally, in Subsection 6.4 we compute the Legendre transform of $`U`$ for the one dimensional zero range process in the presence of external field. As a by product, we show that the macroscopic limit $`N\mathrm{}`$ and $`T\mathrm{}`$ can be interchanged. ## 6.1 A sufficient condition for $`\mathrm{\Phi }=U`$ We consider the case when the matrices $`D(\rho )`$ and $`\chi (\rho )`$ are multiple of the identity, i.e., there are strictly positive scalar functions still denoted by $`D(\rho )`$, $`\chi (\rho )`$, so that $`D(\rho )_{i,j}=D(\rho )\delta _{i,j}`$, $`\chi (\rho )_{i,j}=\chi (\rho )\delta _{i,j}`$, $`i,j=1,\mathrm{},d`$. We denote derivatives with a superscript. Let us first consider the case with no external field, i.e. $`E=0`$, we shall prove that if $$D(\rho )\chi ^{\prime \prime }(\rho )D^{}(\rho )\chi ^{}(\rho )\text{ for any }\rho $$ (6.2) then $`\mathrm{\Phi }=U`$. In this case $`U`$ is necessarily convex. Moreover we show that if $$D(\rho )\chi ^{\prime \prime }(\rho )=D^{}(\rho )\chi ^{}(\rho )\text{ for any }\rho $$ (6.3) then we have $`\mathrm{\Phi }=U`$ for any external field $`E`$. We mention that under the condition (6.3), as shown in \[9, §7\], also the non equilibrium free energy $`F`$ can be computed explicitly and it is a local functional. Condition (6.2) is satisfied for the symmetric simple exclusion process, where $`D=1`$ and $`\chi (\rho )=\rho (1\rho )`$, $`\rho [0,1]`$. Condition (6.3) is satisfied for the zero range model, where $`D(\rho )=\mathrm{\Psi }^{}(\rho )`$ and $`\chi (\rho )=\mathrm{\Psi }(\rho )`$ for some strictly increasing function $`\mathrm{\Psi }:_+_+`$. Condition (6.3) is also satisfied for the non interacting Ginzburg–Landau model, where, $`\rho `$, $`D(\rho )`$ is an arbitrary strictly positive function and $`\chi (\rho )`$ is constant. Let us consider first the case when $`E=0`$ and condition (6.2) holds. In view of (3.9) and (4.4), to prove that $`\mathrm{\Phi }=U`$ it is enough to show that for each $`j=j(t,u)𝒜_{T,J}`$, i.e., such that $`T^1_0^T𝑑tj(t)=J`$, and $`\rho (t)`$ such that $`_t\rho (t)+j(t)=0`$ we have $$\frac{1}{T}_{[0,T]}(j)=\frac{1}{T}_0^T𝑑t𝒰(\rho (t),j(t))U(J)$$ (6.4) Instead of $`\rho `$ we introduce a new variable $`\alpha `$ so that $`\alpha =d(\rho ):=^\rho 𝑑\rho ^{}D(\rho ^{})`$. Condition (6.2) is then equivalent to the concavity of the function $`X(\alpha ):=\chi \left(d^1(\alpha )\right)`$ where $`d^1`$ is the inverse function of $`d`$. We introduce the functional $$𝒱(\alpha ,j):=𝒰(d^1(\alpha ),j)=\frac{1}{2}j+\frac{1}{2}\alpha ,\frac{1}{X(\alpha )}\left[j+\frac{1}{2}\alpha \right]$$ where we used (3.6). We claim that the functional $`𝒱`$ is jointly convex in $`(\alpha ,j)`$. Let us first show that this implies the lower bound (6.4). We have $$\begin{array}{c}\frac{1}{T}_0^T𝑑t𝒰(\rho (t),j(t))=\frac{1}{T}_0^T𝑑t𝒱(\alpha (t),j(t))\hfill \\ 𝒱(\frac{1}{T}_0^T𝑑t\alpha (t),\frac{1}{T}_0^T𝑑tj(t))\underset{\alpha }{inf}𝒱(\alpha ,J)=\underset{\rho }{inf}𝒰(\rho ,J)=U(J)\hfill \end{array}$$ in which we used the convexity of $`𝒱`$ in the second step and $`j𝒜_{T,J}`$ in the third. To prove that $`𝒱`$ is jointly convex we write $$𝒱(\alpha ,j)=\underset{a}{sup}𝒱_a(\alpha ,j),𝒱_a(\alpha ,j):=j+\frac{1}{2}\alpha ,a\frac{1}{2}a,X(\alpha )a$$ and the supremum is taken over all smooth vector fields $`a=a(u)`$ on $`\mathrm{\Lambda }`$. Since $`X(\alpha )`$ is concave, for each fixed $`a`$ the functional $`𝒱_a`$ is jointly convex. The claim follows. In the case with non vanishing $`E`$, we can use the same argument, but the functional $`𝒱`$ is given by $$𝒱(\alpha ,j):=𝒰(d^1(\alpha ),j)=\frac{1}{2}j+\frac{1}{2}\alpha X(\alpha )E,\frac{1}{X(\alpha )}\left[j+\frac{1}{2}\alpha X(\alpha )E\right]$$ Condition (6.3) is equivalent to $`X^{\prime \prime }(\alpha )=0`$; in this case we can easily show, as before, that $`𝒱`$ is jointly convex. ## 6.2 Periodic boundary conditions In this subsection we consider the case when $`\mathrm{\Lambda }=𝕋`$, the one dimensional torus of side length one and constant external field $`E`$. If there is no external field, $`E=0`$, then it is an equilibrium model; non equilibrium if $`E0`$. In this case we have the possibility of constructing a space time path $`(\rho (t,u),j(t,u))`$ of density and current in the form of a travelling wave for the variational problem (4.2) defining the functional $`\mathrm{\Phi }`$. We shall see that, under some assumptions on the transport coefficient $`D(\rho ),\chi (\rho )`$, this strategy for large $`J`$ is more convenient than taking a density path $`\rho `$ constant in time, so that $`\mathrm{\Phi }(J)<U(J)`$. In the context of periodic boundary conditions, the proof that $`\mathrm{\Phi }=U`$ presented in the previous section applies if $`D`$ is constant. In other words, we have that $`\mathrm{\Phi }(J)=U(J)`$ for all $`J`$ provided $`D`$ is constant, $`\chi `$ is concave and $`E=0`$. Let $`_m(𝕋)`$ the convex set of positive functions $`\rho `$ on the torus $`𝕋`$ such that $`_0^1𝑑u\rho (u)=m`$; we call $`m`$ the mass of $`\rho `$. In this context, $$U(J)=\underset{\rho }{inf}\frac{1}{2}_0^1𝑑u\frac{\{JJ(\rho )\}^2}{\chi (\rho (u))},$$ where the infimum is carried over $`_m(𝕋)`$ and $`J(\rho )`$ was defined in (3.6). For each $`v`$, let $`\mathrm{\Psi }_v:_+`$ be defined by $$\mathrm{\Psi }_v(J)=\underset{\rho }{inf}\frac{1}{2}_0^1𝑑u\frac{\{J+v[\rho (u)m]J(\rho )\}^2}{\chi (\rho (u))},$$ (6.5) where the infimum is carried over $`_m(𝕋)`$. We claim that $$\mathrm{\Phi }\mathrm{\Psi }_v$$ (6.6) for each $`v`$. Indeed, consider a profile $`\rho _0`$ in $`_m(𝕋)`$. Let $`T=v^1`$ and set $`\rho (t,u)=\rho _0(uvt)`$, $`j(t,u)=J+v[\rho _0(utv)m]`$ in the time interval $`[0,T]`$. An elementary computation shows that the continuity equation holds and that the time average over the time interval $`[0,T]`$ of $`j(,u)`$ is equal $`J`$. In particular, $$\mathrm{\Phi }(J)\frac{1}{T}_0^T𝑑t𝒰(\rho (t),j(t)).$$ On the other hand, it is easy to show by periodicity that the right hand side is equal to $$\frac{1}{2}_0^1𝑑u\frac{\{J+v[\rho _0(u)m]J(\rho _0)\}^2}{\chi (\rho _0(u))}.$$ Optimizing over the profile $`\rho _0`$, we conclude the proof of (6.6). Fix a mass $`m`$, an external field $`E`$ and a current $`J`$. If $`J^2/\chi +E^2\chi `$ is a convex function then $$U(J)=\frac{1}{2}\frac{\{JE\chi (m)\}^2}{\chi (m)}$$ (6.7) and the optimal profile for the variational problem defining $`U(J)`$ is the constant profile $`\rho (u)=m`$. In particular if $`1/\chi `$ and $`E^2\chi `$ are convex functions then (6.7) holds for any $`J`$ so that $`U`$ is trivially convex. Indeed, fix a mass $`m`$, a current $`J`$ and an external field $`E`$. For any profile $`\rho `$ in $`_m(𝕋)`$, $$_0^1𝑑u\frac{\{JJ(\rho )\}^2}{\chi (\rho )}=_0^1𝑑u\frac{\{JE\chi (\rho )\}^2}{\chi (\rho )}+_0^1𝑑u\frac{[(1/2)d(\rho )]^2}{\chi (\rho )}$$ because the cross term vanishes upon integration Here $`d(\rho )`$ is the function introduced in the previous Subsection. We have $$_0^1𝑑u\frac{\{JE\chi (\rho )\}^2}{\chi (\rho )}=_0^1𝑑u\left[\frac{J^2}{\chi (\rho )}+E^2\chi (\rho )\right]2EJ\frac{J^2}{\chi (m)}+E^2\chi (m)2EJ$$ where we used Jensen inequality, the convexity of $`J^2/\chi +E^2\chi `$, and $`\rho _m(𝕋)`$. Therefore, for all profiles $`\rho `$ in $`_m(𝕋)`$, $$_0^1du\frac{[JJ(\rho )]^2}{\chi (\rho )}\frac{\{JE\chi (m)\}^2}{\chi (m)}$$ Since the cost of the constant profile $`\rho (u)=m`$ is $`(1/2)\{JE\chi (m)\}^2/\chi (m)`$, (6.7) is proven. For the KMP model we have $`D(\rho )=1`$ and $`\chi (\rho )=\rho ^2`$. Since $`\chi `$ and $`1/\chi `$ are convex functions, the assumptions for (6.7) are satisfied for any external field $`E`$. For the simple exclusion process we have $`D(\rho )=1`$ and $`\chi (\rho )=\rho (1\rho )`$. In the case of no external field, $`E=0`$, since $`\chi `$ is concave and $`\chi ^1`$ is convex, it satisfies both the hypotheses for $`\mathrm{\Phi }=U`$ and the ones for (6.7); hence $`\mathrm{\Phi }(J)=(1/2)J^2/m(1m)`$. If $`E0`$ the assumptions for (6.7) holds only if $`|E/J|`$ is small enough. Fix a mass $`m`$, $`e`$ and take the external field $`E=eJ`$. Assume that $$(1e^2\chi (m)^2)\chi ^{\prime \prime }(m)>\mathrm{\hspace{0.33em}0}$$ (6.8) We claim that there exists $`w`$ in $``$ such that $$\underset{|J|\mathrm{}}{lim\; sup}\frac{\mathrm{\Psi }_{wJ}(J)}{J^2}<\frac{\{1e\chi (m)\}^2}{2\chi (m)}$$ (6.9) where $`\mathrm{\Psi }_v`$ has been defined in (6.5). Fix a mass $`m`$, a current $`J`$, an external field $`E=eJ`$ and take $`v=wJ`$. For $`_m(𝕋)`$, by expanding the square we get that $`{\displaystyle _0^1}𝑑u{\displaystyle \frac{\{J+wJ[\rho m]+(1/2)d(\rho )E\chi (\rho )\}^2}{\chi (\rho )}}`$ (6.10) $`=J^2{\displaystyle _0^1}du{\displaystyle \frac{\{1+w[\rho m]e\chi (\rho )\}^2}{\chi (\rho )}}+{\displaystyle \frac{1}{4}}{\displaystyle _0^1}du{\displaystyle \frac{[d(\rho )]^2}{\chi (\rho )}}`$ because the cross term vanishes. Expand the square on the first integral. Let $`F(r)=F_{w,m}(r)`$ be the smooth function defined by $$F(r)=\frac{\{1+w[rm]\}^2}{\chi (r)}2e+e^2\chi (r)$$ An elementary computation shows that $$F^{\prime \prime }(m)=\frac{1}{\chi (m)^3}\left\{2\chi (m)^2w^24\chi (m)\chi ^{}(m)w+2\chi ^{}(m)^2\chi (m)\chi ^{\prime \prime }(m)+e^2\chi ^{\prime \prime }(m)\chi (m)^3\right\}.$$ Let $`w=\chi ^{}(m)/\chi (m)`$. For this choice $`F^{\prime \prime }(m)<0`$. In particular, we can choose a non constant profile $`\rho (u)`$ in $`_m(𝕋)`$ close to $`m`$ such that $`F^{\prime \prime }(\rho (u))<0`$ for every $`u`$. Hence, by Jensen inequality, the coefficient of $`J^2`$ in (6.10) is strictly less than $$\frac{\{1e\chi (m)\}^2}{\chi (m)}$$ The statement follows. Fix $`e`$, assume that $`1/\chi +e^2\chi `$ is convex and (6.8) holds. Take $`E=eJ`$. Then, by (6.6), (6.7) and (6.9) we have $`\mathrm{\Phi }(J)<U(J)`$ for all sufficiently large currents $`J`$. The KMP model satifies the above requirements if $`|e|<1/m^2`$. The exclusion process with external field $`E`$ satisfies (6.8) for $`|E|>|J|/m(1m)`$ but, for these values of $`E`$ and $`J`$ the function $`1/\chi +e^2\chi `$ is not convex so we do not know whether $`U`$ is given by (6.7). By (6.9) we have that, for large current $`J`$, there exists a travelling wave whose cost is strictly less than the one of the constant profile $`\rho (u)=m`$. This is not enough to prove the strict inequality $`\mathrm{\Phi }(J)<U(J)`$. We conclude by giving, for the KMP process with no external field, a description of the mechanism for the strict inequality $`\mathrm{\Phi }<U`$ in terms of the power necessary, according to (A.10), to substain a time average current $`J`$. To get $`U(J)=(1/2)J^2/m^2`$ we switch on a constant external field equal to $`J/m^2`$ which provides exactly the power $`U(J)`$. On the other hand we can impose a time average current $`J`$ by imposing a space time dependent external field of the type $`F(uvt)`$; the corresponding density and current paths are then travelling waves. By exploiting the convexity of $`\rho ^2`$ we have shown that the second strategy, at least for $`J`$ large, requires less power. ## 6.3 An example with non convex $`U`$ We discuss here a special choice of the macroscopic transport coefficients $`D`$ and $`\chi `$ for which the functional $`U`$ defined in (4.5) is not convex. In particular the upper bound (4.7) with the convex envelope differs from (4.6). We take $`d=1`$, $`\mathrm{\Lambda }=(0,1)`$, $`E=0`$, $`D(\rho )=1`$, and $`\chi (\rho )`$ a smooth function with $`\chi (0)=\chi (1)=0`$ (accordingly the density satisfies $`0\rho 1`$) such that there exist $`0<A<B<1`$, $`\mathrm{}`$ for which $`\chi (\rho )=e^\mathrm{}\rho `$ if $`A\rho B`$. Furthermore we take the equilibrium boundary conditions $`\rho (0)=\rho (1)=\overline{\rho }=(A+B)/2`$. We show that, for a suitable choice of the parameters $`A,B,\mathrm{}`$, there are $`J_1<J_2`$ so that $`U^{\prime \prime }(J)<0`$ for any $`J(J_1,J_2)`$. Although we did not construct explicitly a microscopic lattice gas model in which the macroscopic transport coefficients meet the above requirements, we believe it would be possible to exhibit a model which has the same qualitative behavior. For $`d=1`$, $`D=1`$, $`E=0`$, and $`J`$, the Euler–Lagrange equation for the variational problem (4.5) defining the functional $`U(J)`$ is $$\{\begin{array}{c}\frac{1}{2}\rho ^{\prime \prime }(u)=\frac{\chi ^{}(\rho (u))}{\chi (\rho (u))}\left[J^2\frac{1}{4}\rho ^{}(u)^2\right],u(0,1)\hfill \\ \rho (0)=\rho _0,\rho (1)=\rho _1\hfill \end{array}$$ (6.11) For the above choice of $`\chi `$ and of the boundary conditions, provided $`4|J|BA`$, a solution of (6.11) is given by $$\widehat{\rho }_J(u)=\overline{\rho }+\frac{2}{\mathrm{}}\mathrm{log}\frac{\mathrm{cosh}[J\mathrm{}(u1/2)]}{\mathrm{cosh}[J\mathrm{}/2]}$$ (6.12) as can be easily verified. Note indeed that $`A\widehat{\rho }_JB`$ since we assumed $`4|J|BA`$. We shall later prove that, under the above conditions, (6.12) is the unique solution of the boundary value problem (6.11). A simple computation then gives $$U(J)=\frac{2e^{\mathrm{}\overline{\rho }}}{\mathrm{}^2}\frac{J\mathrm{}}{2}\mathrm{tanh}\frac{J\mathrm{}}{2}$$ Let $`F(z):=z\mathrm{tanh}z`$ and $`z^{}`$ be the unique positive root of $`z^1=\mathrm{tanh}z`$. Then $`F^{\prime \prime }(z)=2(1\mathrm{tanh}^2z)(1z\mathrm{tanh}z)<0`$ for $`z>z^{}`$. Hence $`U^{\prime \prime }(J)<0`$ if $`|J|(2z^{}/\mathrm{},(BA)/4)`$. This interval is not empty provided $`\mathrm{}`$ is chosen large enough. To show that (6.12) is the unique solution of the boundary value problem (6.11), let us first prove that, given $`J0`$, any solution of (6.11) satisfies the *a priori bound* $`|\rho ^{}|2|J|`$. Since $`\rho (0)=\rho (1)`$, we can exclude the possibility that $`|\rho ^{}(u)|2|J|`$ for every $`u[0,1]`$. By the continuity of $`\rho ^{}`$, it is therefore enough to prove that $`|\rho ^{}(u)|2|J|`$ for every $`u[0,1]`$. Suppose conversely that there exists $`u^{}[0,1]`$ such that $`\rho ^{}(u^{})=2J`$. Then, by the uniqueness of the Cauchy problem $`\frac{1}{2}\rho ^{\prime \prime }=\frac{\chi ^{}(\rho )}{\chi (\rho )}\left[J^2\frac{1}{4}\rho ^2\right]`$, $`\rho (u^{})=\rho ^{}`$, $`\rho ^{}(u^{})=2J`$, we would get that the solution of (6.11) is $`\rho (u)=\rho (u^{})+2J(uu^{})`$. Since this function does not satisfy the boundary conditions in (6.11) we find the desired contradiction. Since $`4|J|BA`$, the *a priori bound* $`|\rho ^{}|2|J|`$ implies that any solution $`\rho `$ of (6.11) satisfies $`A\rho B`$. For such values we have that $`\chi ^{}(\rho )/\chi (\rho )=\mathrm{}`$. Uniqueness of the solution to (6.11) can then be easily proven by explicit computations. ## 6.4 Zero range processes In this section we consider the so-called one-dimensional zero-range processes which models a non-linear diffusion of lattice gases under constant external field $`E`$. The model is described by positive integer-valued variables $`\eta _x`$ representing the number of particles at site $`x`$. The particles jump with rates $`(1/2)g(\eta _x))\mathrm{exp}\{E/N\}`$ to right, $`(1/2)g(\eta _x))`$ $`\mathrm{exp}\{E/N\}`$ to the left, respectively. The function $`g(k)`$ is such that $`g(k+1)g(k)a`$ for some $`a>0`$ and $`g(0)=0`$. We assume that our system interacts with particle reservoirs at the sites $`0`$ and $`N`$ whose activity is given by $`\phi _0`$, $`\phi _1`$. The generator of this Markov process is given by (2.3) with $`\mathrm{\Lambda }_N=\{1,\mathrm{},N\}`$ and $`c_{x,x+1}(\eta )=g(\eta _x)e^{E/N},c_{x+1,x}(\eta )=g(\eta _{x+1})e^{E/N}`$ for $`1xN1`$. Moreover, at the boundary, $`c_{0,1}(\eta )=\phi _0e^{E/N},c_{1,0}(\eta )=g(\eta _1)e^{E/N},`$ $`c_{N,N+1}(\eta )=g(\eta _N)e^{E/N},c_{N+1,N}(\eta )=\phi _1e^{E/N}.`$ Let $`V(u)=Eu`$ and let $`\phi _N(x)`$ be the solution of $$e^V\mathrm{\Delta }_N\frac{\phi _N}{e^V}=\frac{\phi _N}{e^V}\mathrm{\Delta }_Ne^V$$ for $`1xN`$ and with boundary condition $`\phi _N(0)=\phi _0`$, $`\phi _N(N+1)=\phi _1`$. Here $`\mathrm{\Delta }_N`$ stands for the discrete Laplacian. The invariant measure $`\mu _N`$ is the grand–canonical measure $`\mu _N=_{x\mathrm{\Lambda }_N}\mu _{x,N}`$ obtained by taking the product of the marginal distributions $$\mu _{x,N}(\eta _x=k)=\frac{1}{Z(\phi _N(x))}\frac{\phi _N(x)^k}{g(1)\mathrm{}g(k)}$$ where $`Z(\phi )=_{k0}\phi ^k/[g(1)\mathrm{}g(k)]`$ is the normalization constant. Let $`R(\phi )=\phi Z^{}(\phi )/`$ $`Z(\phi )`$ and denote by $`\mathrm{\Psi }`$ its inverse function. For this process, the hydrodynamic equation (1.2) and the large deviations principle (1.4) can be obtained with $`D(\rho )=\mathrm{\Psi }^{}(\rho )`$, $`\chi (\rho )=\mathrm{\Psi }(\rho )`$, see . Moreover, $`F_0^{}(\rho )=\mathrm{log}\mathrm{\Psi }(\rho )`$. Let first show how, for this model, it is possible to solve explicitly the variational problem (4.11) for the Laplace transform of the total current. We note that the case $`E=0`$ has already been solved in . Since we are in one space dimension, the condition $`\lambda =0`$ simply states that $`\lambda `$ is a constant. Moreover we have $`J(\rho )=(1/2)\mathrm{\Psi }^{}(\rho )\rho ^{}+\mathrm{\Psi }(\rho )E`$, where hereafter the apices denotes differentiation w.r.t. the macroscopic variable $`u`$. Note that condition (6.3) holds so that $`\mathrm{\Phi }=U`$. Changing variables in (4.11) by introducing $`\phi (u)=\mathrm{\Psi }(\rho (u))`$, $`u[0,1]`$; we get $$U^{}(\lambda )=\frac{1}{2}\underset{\phi }{sup}_0^1𝑑u\left\{\phi (u)\psi ^{}(u)^2+\lambda ^2\phi (u)\lambda \phi ^{}(u)+2\lambda E\phi (u)\right\}$$ where the supremum is over all positive $`\phi `$ such that $`\phi (0)=\phi _0`$, $`\phi (1)=\phi _1`$ and $`\psi `$ solves $$2\left(\phi (u)\psi ^{}(u)\right)^{}=\left(\phi ^{}(u)+2[E+\lambda ]\phi (u)\right)^{}$$ with boundary conditions $`\psi (0)=\psi (1)=0`$. The solution is $$\psi (u)=\frac{1}{2}\mathrm{log}\frac{\phi (u)}{\phi _0}+(E+\lambda )u+A_0^u𝑑v\frac{1}{\phi (v)}$$ where $$A=\left\{\frac{1}{2}\mathrm{log}\frac{\phi _1}{\phi _0}(E+\lambda )\right\}\left\{_0^1𝑑u\frac{1}{\phi (u)}\right\}^1$$ After elementary manipulations, the variational problem for $`U^{}(\lambda )`$ becomes $$\begin{array}{c}U^{}(\lambda )=\frac{1}{2}\underset{\phi }{sup}\{\{\frac{1}{2}\mathrm{log}\frac{\phi _1}{\phi _0}+E+\lambda \}^2\left\{_0^1du\frac{1}{\phi (u)}\right\}^1\hfill \\ \frac{1}{4}_0^1du\frac{\phi ^{}(u)^2}{\phi (u)}E^2_0^1du\phi (u)+E(\phi _1\phi _0)\}\hfill \end{array}$$ The associated extremality condition, which determines the optimal profile, is $$2\phi ^{\prime \prime }(u)\phi (u)\left(\phi ^{}(u)\right)^24E^2\phi (u)^2=4\left(\frac{1}{2}\mathrm{log}\frac{\phi _1}{\phi _0}+E+\lambda \right)^2\left[_0^1𝑑v\frac{1}{\phi (v)}\right]^2$$ (6.13) with the boundary condition $`\phi (0)=\phi _0`$, $`\phi (1)=\phi _1`$. In the case $`E=0`$ it is not difficult to check that the solution of (6.13) is $$\phi (u)=C\left(u+\frac{e^\lambda }{1e^\lambda }\right)\left(u\frac{\phi _0e^\lambda }{\phi _0e^\lambda \phi _1}\right)$$ where $`C=\left(1e^\lambda \right)\left(\phi _0e^\lambda \phi _1\right)`$. We then get $$U^{}(\lambda )=\frac{1}{4}[\phi ^{}(1)\phi ^{}(0)]=\frac{1}{2}\left(1e^\lambda \right)\left(\phi _0e^\lambda \phi _1\right)$$ In the case $`E0`$, the solution of (6.13) is instead given by $$\phi (u)=C\left(e^{2Eu}a\right)\left(e^{2Eu}b\right)$$ where $$a=\frac{\phi _0e^{2E+\lambda }\phi _1}{\phi _0e^\lambda \phi _1}b=\frac{1e^{\lambda 2E}}{1e^\lambda }$$ and $$C=\frac{(1e^\lambda )(\phi _0e^\lambda \phi _1)}{(e^{2E}1)(1e^{2E})}$$ Notice that this solution converges, as $`E0`$, to the solution with no external field. Plugging this solution into the variational formula for $`U^{}`$, we get that $$U^{}(\lambda )=\frac{1}{4}[\phi ^{}(1)\phi ^{}(0)]+\frac{1}{2}[\phi _1\phi _0]=E\left\{\frac{\phi _0}{1e^{2E}}(e^\lambda 1)+\frac{\phi _1}{e^{2E}1}(e^\lambda 1)\right\}.$$ (6.14) We conclude this section showing that we may invert the order of limits in (4.9) for zero range models. For $`0x,yN+1`$, $`|xy|=1`$, recall that we denote by $`𝒩_t^{x,y}`$ the total number of jumps from $`x`$ to $`y`$ in the time interval $`[0,t]`$. For $`0xN`$, let $`Q_t^{x,x+1}=𝒩_t^{x,x+1}𝒩_t^{x+1,x}`$ be the total current over the bond $`(x,x+1)`$. Note that we are including the boundary bonds. Consider the limit as microscopic time $`t`$ goes to infinity of the Laplace transform of the total current: $$e_N(\lambda )=\frac{1}{N}\underset{t\mathrm{}}{lim}\frac{1}{t}\mathrm{log}𝔼_{\eta ^N}^N\left[\mathrm{exp}\left\{\lambda N^1\underset{x=0}{\overset{N}{}}Q_t^{x,x+1}\right\}\right]$$ and notice that $`f_N`$ given by (5.11) is related to the Legendre transform of $`e_N`$ by $$f_N^{}(\lambda )=e_N\left(\lambda \left\{2E\mathrm{log}(\phi _1/\phi _0)\right\}\right).$$ Notice furthermore that this expression does not depend on the initial condition $`\eta ^N`$ by ergodicity. Since two currents $`Q_t^{x,x+1}`$ $`Q_t^{y,y+1}`$ differ only by surface terms, in the asymptotic $`t\mathrm{}`$, we may replace all currents by $`Q_t^{0,1}`$ and obtain that $$e_N(\lambda )=\frac{1}{N}\underset{t\mathrm{}}{lim}\frac{1}{t}\mathrm{log}𝔼_\mu \left[e^{\lambda Q_t^{0,1}}\right].$$ To compute the previous limit, we represent the zero range process in terms of interacting random walks. Let $`N_0`$ be the total number of particles at time $`0`$: $`N_0=_x\eta _x(0)`$. We start labeling these particles. New particles entering the system at the boundary get new labels in an increasing order. Denote by $`X^i(t)`$ the position at time $`t`$ of the $`i`$-th particle. $`X^1`$ performs a weakly asymmetric random walk on $`\mathrm{\Lambda }_N`$ with absorption at the boundary and mean $`g(1)`$ exponential waiting times. $`X^2`$ does the same but its clock rates are affected by $`X^1`$. If they occupy different sites, the $`X^2`$-exponential has rate $`g(1)`$, while if both occupy the same site, its exponential clock has rate $`g(2)g(1)`$ and so on. We need for this construction the function $`g`$ to be increasing. Moreover the condition $`g(k+1)g(k)a>0`$ guarantees that these random walks will hit the boundary with probability one. Let $`w_i`$ (resp. $`u_i`$) the indicator function of the event that the $`i`$-th particle created at the left (resp. right) boundary is absorbed at $`N+1`$ (resp. $`0`$). Denote by $`N_\pm (t)`$ a Poisson process of rate $`\phi _0/2`$, $`\phi _1/2`$ which represents the entrance of particles at either boundary. With this notation, up to negligible terms in the limit $`t\mathrm{}`$, $$Q_t^{0,1}=\underset{i=1}{\overset{N_{}(t)}{}}w_i\underset{i=1}{\overset{N_+(t)}{}}u_i.$$ Since the random variables $`w_i`$, $`u_j`$ are independent, elementary computation shows that $$e_N(\lambda )=\frac{1}{N}\left\{(\phi _0/2)p_N\{e^\lambda 1\}+(\phi _1/2)q_N\{e^\lambda 1\}\right\},$$ where $`p_N=P[w_i=1]`$ (resp. $`1q_N`$) is the probability that a random walk, absorbed in $`0`$ and $`N+1`$, with transition probability $`p(x,x+1)=e^{2E/N}/(e^{2E/N}+1)=1p(x,x1)`$ starting from $`1`$ (resp. $`N`$) is absorbed in $`N+1`$ (resp. $`0`$). These probabilities can be explicitly computed. As $`N\mathrm{}`$, we get that $`\underset{N\mathrm{}}{lim}e_N(\lambda )=E\left\{{\displaystyle \frac{\phi _0}{1e^{2E}}}\left(e^\lambda 1\right)+{\displaystyle \frac{\phi _1}{e^{2E}1}}\left(e^\lambda 1\right)\right\}`$ which agrees with (6.14). A. Supplement to Section 1.3. We present here a derivation, at the heuristic level, of the law of large numbers and the large deviations, as $`N\mathrm{}`$, for the empirical density and the empirical current. Recall the notation introduced in Section 1.3. We have seen there that to prove the law of large numbers for the empirical measure and the current, we need to express the limit current $`J`$ in terms of the density $`\rho `$. In the context of stochastic lattice gases this is done by assuming a *local equilibrium state*. Roughly speaking, this means that in a large microscopic region $`\mathrm{\Delta }`$ around $`u`$, still infinitesimal macroscopically, the system has relaxed to the Gibbs state (with Hamiltonian $``$) conditioned to $`_{x\mathrm{\Delta }}\eta _x=|\mathrm{\Delta }|\pi ^N(t,u)`$. This assumption, which can be rigorously justified, , allows us to express the empirical current in terms of the empirical density. We next show how this can be done for the so–called *gradient* models. By standard computations in the theory of Markov processes we have that, \[31, Lemma II.2.3\], for a bond $`\{x,x+e_j\}`$, $$Q^{x,x+e_j}(t)=(1/2)N^2_0^t𝑑s\left[c_{x,x+e_j}(\eta _s)c_{x+e_j,x}(\eta _s)\right]+M^{x,x+e_j}(t),$$ where $`M^{x,x+e_j}(t)`$ are martingales with bracket $$M^{x,x+e_i},M^{y,y+e_j}(t)=(1/2)N^2\delta _{x,y}\delta _{i,j}_0^t𝑑s\left[c_{x,x+e_i}(\eta (s))+c_{x+e_i,x}(\eta (s))\right].$$ Let $`G`$ be a smooth vector field as in (3.2) vanishing on $`\mathrm{\Lambda }`$. By definition of the martingales $`M^{x,x+e_j}(t)`$, $$𝒥^N,G_T=\frac{1}{2}\frac{1}{N^d}_0^T𝑑t\underset{i=1}{\overset{d}{}}\underset{x}{}G_i(t,x/N)N\left[c_{x,x+e_i}(\eta (t))c_{x+e_i,x}(\eta (t))\right]+_T^N(G),$$ (A.1) where $`_T^N(G)`$ is a martingale term. An easy computation, based on the explicit formula for the quadratic variations of the martingales $`M^{x,x+e_j}(t)`$, shows that $`_T^N(G)`$ vanishes as $`N\mathrm{}`$. We next use definition (2.2) and Taylor expansion to write $$\begin{array}{c}c_{x,x+e_i}(\eta )c_{x+e_i,x}(\eta )\hfill \\ =c_{x,x+e_i}^0(\eta )\left[1+\frac{1}{N}E_i(x/N)\right]c_{x+e_i,x}^0(\eta )\left[1\frac{1}{N}E_i(x/N)\right]+O(1/N^2)\hfill \\ =\left[c_{x,x+e_i}^0(\eta )c_{x+e_i,x}^0(\eta )\right]+\frac{1}{N}\left[c_{x,x+e_i}^0(\eta )+c_{x+e_i,x}^0(\eta )\right]E_i(x/N)+O(1/N^2)\hfill \end{array}$$ (A.2) The *gradient condition*, see \[31, II.2.4\], holds if there exist local functions $`h_0^{(i)}(\eta )`$, $`i=1,\mathrm{},d`$, depending on the configuration $`\eta `$ around $`0`$, so that for any $`i=1,\mathrm{},d`$ $$c_{x,x+e_i}^0(\eta )c_{x+e_i,x}^0(\eta )=h_x^{(i)}(\eta )h_{x+e_i}^{(i)}(\eta )$$ (A.3) where $`h_x^{(i)}`$ is the function $`h_0^{(i)}`$ evaluated on the configuration $`\eta `$ translated by $`x`$. Let us plug the right hand side of (A.2) into (A.1). By the gradient condition (A.3) we can perform a summation by parts on the first term. Note that there are no boundary terms since we assumed $`G`$ to vanish on the boundary. We get, with a negligible error as $`N\mathrm{}`$, $$\begin{array}{c}𝒥^N,G_T\frac{1}{2}\frac{1}{N^d}_0^Tdt\underset{i=1}{\overset{d}{}}\underset{x}{}\{_iG_i(t,x/N)h_x^{(i)}(\eta )\hfill \\ +G_i(t,x/N)[c_{x,x+e_i}^0(\eta (t))+c_{x+e_i,x}^0(\eta (t))]E_i(x/N)\}\hfill \end{array}$$ (A.4) Recall the definition of the functions $`d^{(i)}`$, $`\chi ^{(i)}`$ introduced in (3.5). By the local equilibrium assumption mentioned above and the equivalence of ensembles from (A.4) we get $$𝒥^N,G_T\underset{i=1}{\overset{d}{}}_0^T𝑑t_\mathrm{\Lambda }𝑑u\left\{\frac{1}{2}_iG_i(t,u)d^{(i)}\left(\pi ^N(t,u)\right)+G_i(t,u)\chi ^{(i)}\left(\pi ^N(t,u)\right)E_i(u)\right\}$$ (A.5) Taking the limit $`N\mathrm{}`$, the empirical density $`\pi ^N(t,u)`$ converges to $`\rho (t,u)`$, whereas the empirical current $`𝒥^N(t,u)`$ converges to a vector field $`J(t,u)`$. Equation (A.5) then implies $$J(\rho )=\frac{1}{2}D(\rho )\rho +\chi (\rho )E$$ (A.6) where $`D`$ and $`\chi `$ are $`d\times d`$ diagonal matrices with entries $`D_{ii}(\rho )=\frac{d}{d\rho }d^{(i)}(\rho )`$ and $`\chi _{ii}(\rho )=\chi ^{(i)}(\rho )`$. We now turn to a heuristic derivation of the large deviations principle (3.8)–(3.9) for the current. Recall the statement and the notation introduced in Section 1.3. In order to make the trajectory $`j`$ typical, we introduce an extra weak time dependent external field $`F=(F_1,\mathrm{},F_d)`$ by perturbing the rates as in Section 1.3, namely $$c_{x,x+e_i}^F(\eta )=c_{x,x+e_i}(\eta )e^{N^1F_i(t,x/N)},c_{x+e_i,x}^F(\eta )=c_{x+e_i,x}(\eta )e^{N^1F_i(t,x/N)}.$$ (A.7) We denote by $`_{\eta ^N}^{N,F}`$ the probability distribution of the perturbed process. Since these rates $`c^F`$ are the same as the rates of the original process with $`E`$ replaced by $`E+F`$ (cf. (2.2)), we have the following law of large numbers: $$\underset{N\mathrm{}}{lim}_{\eta ^N}^{N,F}\left(𝒥^Nj\right)=1$$ where $$j=J(\rho )+\chi (\rho )F=\frac{1}{2}D(\rho )\rho +\chi (\rho )(E+F)$$ (A.8) and $`\rho `$ satisfies the continuity equation (3.10). We now read this equation in the opposite direction: given the trajectory $`j`$ we first solve (3.10) to get $`\rho `$, then we determine the external field $`F`$ which makes $`j`$ the typical behavior, namely $$F=\chi (\rho )^1\left(j+\frac{1}{2}D(\rho )\rho \right)E$$ (A.9) By writing the original process in terms of the perturbed one we have $$_{\eta ^N}^N\left(𝒥^N(t,u)j(t,u),(t,u)[0,T]\times \mathrm{\Lambda }\right)=_{\eta ^N}^{N,F}\left(\frac{d_{\eta ^N}^N}{d_{\eta ^N}^{N,F}}\mathrm{\hspace{0.17em}1}\mathrm{I}_{\{𝒥^Nj\}}\right)$$ The large deviation principle (3.8)–(3.9) will follow, recalling (A.9), once we compute the Radon–Nikodym derivative and show that on the event $`\{𝒥^Nj\}`$ we have, with a negligible error if $`N\mathrm{}`$, $$\mathrm{log}\frac{d_{\eta ^N}^N}{d_{\eta ^N}^{N,F}}=\mathrm{log}\frac{d_{\eta ^N}^{N,F}}{d_{\eta ^N}^N}N^d\frac{1}{2}_0^T𝑑tF,\chi (\rho )F$$ (A.10) This equation can be interpreted, in analogy to the classical Ohm’s law, as the total work done in the time interval $`[0,T]`$ by the external field $`F`$. We shall need some basic tool from the general theory of jump Markov processes that we briefly recall, see e.g. \[20, Appendix A1\] or \[5, Appendix A\]. Let $`\mathrm{\Omega }`$ be a countable set and consider a continuous time jump Markov process $`X_t`$ on the state space $`\mathrm{\Omega }`$ with generator given by $$Lf(\eta )=\underset{\eta ^{}\mathrm{\Omega }}{}\lambda (\eta )p(\eta ,\eta ^{})\left[f(\eta ^{})f(\eta )\right]$$ (A.11) where the rate $`\lambda `$ is a positive function on $`\mathrm{\Omega }`$ and $`p(\eta ,\eta ^{})`$ is a transition probability. We consider also another process $`X_t^F`$ of the same type with time dependent rate $`\lambda ^F(\eta ,t)`$ and transition probability $`p^F(\eta ,\eta ^{};t)`$. Then, denoting by $`_{\eta _0}`$ and $`_{\eta _0}^F`$ the distribution of the two processes with initial condition $`\eta _0`$ we have $$\begin{array}{c}\frac{d_{\eta _0}^F}{d_{\eta _0}}\left(X_t,t[0,T]\right)\hfill \\ =\mathrm{exp}\left\{\underset{i=1}{\overset{n}{}}\mathrm{log}\frac{\lambda ^F(X_{\tau _{i1}},\tau _i)p^F(X_{\tau _{i1}},X_{\tau _i};\tau _i)}{\lambda (X_{\tau _{i1}})p(X_{\tau _{i1}},X_{\tau _i})}_0^T𝑑t\left[\lambda ^F(X_t,t)\lambda (X_t)\right]\right\}\hfill \end{array}$$ (A.12) where $`\tau _0=0`$, $`X_0=\eta _0`$, $`\tau _i`$, $`i=1,\mathrm{},n`$ is the time in which the process jumped from $`X_{\tau _{i1}}`$ to $`X_{\tau _i}`$, and $`n`$ is total number of jumps in the time interval $`[0,T]`$. For the process $`\eta (t)`$ with generator (2.3) we have for $`x`$, $`y`$ in $`^d`$, $$\lambda (\eta )p(\eta ,\sigma ^{x,y}\eta )=(1/2)N^2c_{x,y}(\eta )\lambda (\eta )=(1/2)N^2\underset{x,y}{}c_{x,y}(\eta )$$ For the process $`\eta (t)`$ with rates (A.7) we have $$\lambda ^F(\eta ,t)p^F(\eta ,\sigma ^{x,x+e_i}\eta ,t)=(1/2)N^2c_{x,x+e_i}(\eta )e^{N^1F_i(t,x/N)}$$ and a similar formula for $`\lambda ^F(\eta ,t)p^F(\eta ,\sigma ^{x+e_i,x}\eta ,t)`$ so that $$\lambda ^F(\eta )=(1/2)N^2\underset{i=1}{\overset{d}{}}\underset{x}{}\left\{c_{x,x+e_i}(\eta )e^{N^1F_i(t,x/N)}+c_{x+e_i,x}(\eta )e^{N^1F_i(t,x/N)}\right\}.$$ From (A.12) and the explicit expressions for the rates, we get that $`\mathrm{log}{\displaystyle \frac{d_{\eta ^N}^{N,F}}{d_{\eta ^N}^N}}={\displaystyle \frac{1}{N}}{\displaystyle \underset{j=1}{\overset{d}{}}}{\displaystyle \underset{x}{}}\left\{{\displaystyle \underset{\tau _{x,x+e_j}}{}}F_j(\tau _{x,x+e_j},x/N){\displaystyle \underset{\tau _{x+e_j,x}}{}}F_j(\tau _{x+e_j,x},x/N)\right\}`$ $`{\displaystyle \frac{N^2}{2}}{\displaystyle _0^T}𝑑t\left\{c_{x,x+e_j}(\eta (t))\left(e^{N^1F(t,x/N)}1\right)+c_{x+e_j,x}(\eta (t))\left(e^{N^1F(t,x/N)}1\right)\right\},`$ where $`\tau _{x,y}`$ are the jump times from $`x`$ to $`y`$. Expanding the exponentials and recalling the definition of the empirical current, we may rewrite the previous expression as $$\begin{array}{c}N^d𝒥^N,F_T\frac{N}{2}_0^T𝑑t\underset{x}{}\underset{i=1}{\overset{d}{}}F_i(x/N,t)\left\{c_{x,x+e_i}(\eta (t))c_{x+e_i,x}(\eta (t))\right\}\hfill \\ \frac{1}{2}_0^T𝑑t\underset{x}{}\underset{i=1}{\overset{d}{}}\left\{c_{x,x+e_i}(\eta (t))+c_{x+e_i,x}(\eta (t))\right\}F_i(x/N,t)^2+O(1/N),\hfill \end{array}$$ (A.13) where we let $`c_{x,y}=0`$ if $`x,y\mathrm{\Lambda }_N`$. For gradient models condition (A.3) holds so that we can perform a summation by parts in the second term. Recalling the definition of the diffusion matrix $`D`$, the mobility $`\chi `$ and the local equilibrium assumption, we can express the second term of the right hand side of (A.13) in terms of the empirical density. Since we are assuming $`𝒥^Nj`$, we get that $$\mathrm{log}\frac{d_{\eta ^N}^{N,F}}{d_{\eta ^N}^N}N^d\left\{j,F_T+_0^T𝑑tF,(1/2)D(\rho )\rho \chi (\rho )E\frac{1}{2}_0^T𝑑tF,\chi (\rho )F\right\}.$$ (A.14) which, by the choice of $`F`$ in (A.9), concludes the derivation of (A.10). The rigorous derivation of action functional $``$ requires some difficult estimates. In fact, while in the proof of the hydrodynamic limit it is enough to show that the local equilibrium assumption holds with a negligible error as $`N\mathrm{}`$, in the proof of the large deviations we need such an error to be $`o(e^{CN^d})`$. This can be proven by the so called super exponential estimate, see , which is the key point in the rigorous approach. Recall the dynamical large deviations principle for the density stated in (1.4). The rate functional $``$ is given by $$_{[0,T]}(\rho )=\frac{1}{2}_0^T𝑑tH(t),\chi (\rho (t))H(t)$$ (A.15) where, given the fluctuation $`\rho `$, the external potential $`H=H(t,u)`$ is chosen so that it vanishes at the boundary and $$_t\rho =\left(\frac{1}{2}D(\rho )\rho \chi (\rho )\left[E+H\right]\right)$$ (A.16) which is a Poisson equation for $`H`$. We conclude this Appendix showing how the above result follows directly from the large deviation principle for the current. We fix a path $`\rho =\rho (t,u)`$, $`(t,u)[0,T]\times \mathrm{\Lambda }`$. There are many possible trajectories $`j=j(t,u)`$, differing by divergence free vector fields, such that the continuity equation (3.10) is satisfied. The functional $`_{[0,T]}(\rho )`$ can be obtained by minimizing $`_{[0,T]}(j)`$ among all such paths $`j`$ $$_{[0,T]}(\rho )=\underset{\begin{array}{c}j:\\ j=_t\rho \end{array}}{inf}_{[0,T]}(j)$$ (A.17) To derive the functional (A.15) we show that the infimum above is obtained when the external perturbation $`F`$ introduced in (A.9) is a gradient vector field whose potential $`H`$ solves (A.16). Let $`H`$ be the solution of (A.16) and $`F`$ as in (A.9), we write $$F=H+\stackrel{~}{F}$$ (A.18) By the definition of $`H`$ we get $$H,\chi (\rho )\stackrel{~}{F}=H,j+\left(\frac{1}{2}D(\rho )\rho \chi (\rho )E\chi (\rho )H\right)=0$$ Hence $$_{[0,T]}(j)=\frac{1}{2}_0^T𝑑t\left\{H,\chi (\rho )H+\stackrel{~}{F},\chi (\rho )\stackrel{~}{F}\right\}$$ Therefore the infimum in (A.17) is obtained when $`\stackrel{~}{F}=0`$, so that the functional defined in (A.17) coincides with (A.15). Acknowledgments. It is a pleasure to thank T. Bodineau, B. Derrida, G. Gallavotti, E. Presutti, C. Toninelli, and S.R.S. Varadhan for stimulating discussions. The authors acknowledge the support of PRIN MIUR 2004028108 and 2004015228.
warning/0506/quant-ph0506002.html
ar5iv
text
# Simulation of interaction Hamiltonians by quantum feedback: a comment on the dynamics of information exchange between coupled systems ## 1 Introduction In recent years, there has been rapid progress in the analysis of quantum systems at a fundamental level based on precise measurements and optimized control. Experimentally, optical systems have provided a fertile testing ground for this new understanding of quantum effects due to the availability of highly coherent lasers and sensitive detectors. In the light of these new technologies, the discussion about quantum measurement has obtained a practical relevance that sometimes challenges the seemingly well-defined notions of quantum states and Hamiltonians conveyed by typical introductions to quantum mechanics. A particularly instructive example is the development of quantum feedback theory, which originated from a formal analysis of open system dynamics that was largely motivated by the intention to identify the proper pure state description in the presence of noise, but resulted instead in a new formulation of conditional quantum dynamics that actually highlights the inadequacy of a measurement independent definition of pure states in the dynamics of open systems . By emphasizing the importance of classically available measurement information, the theory of quantum feedback then permitted the identification of interesting parallels between the classical notion of control and its quantum mechanical equivalent , a result that may be of significant practical use. My own interest in quantum feedback originated from studies of quantum noise in lasers, where the classical description of light is surprisingly successful even when the intensities studied are so small that individual photons could be resolved. This observation can be explained in detail by quantum measurement theory: in the presence of a strong laser field, the main effect of the spontaneous emission from a single atom is its interference with the laser light, not the energy contributed by the single atom. It is thus reasonable to analyze the dynamics of single atom emission in terms of homodyne detection of the electromagnetic field. Interestingly, quantum dynamics then corresponds closely to classical electrodynamics, and the effects of quantum noise and of classical dipole radiation can be identified in the measurement statistics. By using feedback, it is therefore possible to eliminate the quantum noise effects and to reduce the effect of spontaneous emission to a quantum nondemolition measurement of the atomic dipole . This method may have interesting applications to the stabilization of atomic quantum states, and it has been pointed out that, in principle, this kind of quantum feedback can indeed stabilize almost any quantum state of a two level atom . Quantum feedback can thus restore some of the classical concepts of dynamics that seem to be lost in the transition from classical systems to quantum systems. In particular, the measurement information used in feedback can always be interpreted as a minimal back-action measurement of a system variable and a conditional unitary evolution of the system . A feedback setup therefore keeps track of some of the observable properties of the system, in addition to the information available on the conditional forces responsible for the deterministic evolution of the system . It may be worthwhile to persue this line of thought a bit further in order to gain a better understanding of the relationship between the unobservable closed system dynamics and the observable (and therefore controllable) dynamics of quantum feedback. Specifically, this approach could shed some light on the quantum mechanical interaction dynamics between two coupled systems by providing a description of interaction that combines the classical notion of deterministic conditional forces with the quantum notion of the interaction Hamiltonian. In the following, I will therefore describe systems where the interaction is realized entirely by quantum feedback. According to standard quantum feedback theory, the dynamics is then described by an effective interaction Hamiltonian and a (seemingly separate) superoperator describing the decoherence associated with the measurement interaction. The feedback system can thus simulate a genuine quantum interaction. By itself, the effective Hamiltonian would entangle the interacting system, but in the context of the feedback setup, the quantum state must remain separable because the quantum operations on the individual systems are in fact local. In the terminology of quantum information theory , the interaction has been realized by local operations and classical communication between the systems, whereas the Hamiltonian interaction itself corresponds to quantum communication between the systems. The result of this analysis shows that the decoherence term is not really separable from the Hamiltonian dynamics, but describes the quantum noise required to reduce the exchange of quantum information between the systems to zero. The transition from a purely classical exchange of traceable information to the more intimate entanglement generating quantum interaction is therefore a quantitative one based on the precise relation between the noise levels and the coupling constants representing the forces acting between the systems. This observation is consistent with recent results on the robustness of quantum gate operations against noise and may therefore have interesting implications for the evaluation of experimental quantum devices. The quantum feedback analysis of interactions between quantum systems thus reveals an interesting connection between basic concepts of quantum information and physical interactions. Moreover, the qualitative correspondence between the feedback dynamics and the quantum interaction described by the Hamiltonian indicates that the interpretation of interactions in terms of conditional forces acting between systems might have its applications even in the context of genuine quantum interactions. The rest of the paper is organized as follows. In section 2, the possibility of modifying the Hamiltonian dynamics by quantum feedback is reviewed. In section 3, these results are applied to a pair of non-interacting systems to generate an effective interaction Hamiltonian. In section 4, the results are applied to the analysis of noisy interactions to derive a criterion for the separability of the interaction dynamics. In section 5, the results are illustrated for the case of a feedback induced interaction between a pair of optical cavity modes. It is shown that the feedback analysis provides exact uncertainty limits for the separability of the two mode squeezing interaction. In section 6, the implications of separability for the information exchange between interacting quantum systems is discussed. The results are summarized in 7. ## 2 Quantum feedback and effective Hamiltonians First, it may be useful to review some of the central results of continuous quantum feedback theory in terms of the relation between measurement information and conditional dynamics. For this purpose, let us consider the dynamics of a quantum system coupled to the environment in such a way that some observable property $`\widehat{X}`$ causes the emission of a corresponding signal in a quantum fluctuating field propagating away from the system to a detector setup. The effects of the environment on the system state then causes dephasing between eigenstates of $`\widehat{X}`$. If this is the only relevant dynamics of the system, the evolution of the density matrix $`\widehat{\rho }`$ can be written as $$\frac{d}{dt}\widehat{\rho }=\gamma \left(\widehat{X}\widehat{\rho }\widehat{X}\frac{1}{2}\widehat{X}^2\widehat{\rho }\frac{1}{2}\widehat{\rho }\widehat{X}^2\right),$$ (1) where $`\gamma `$ is the coupling rate determining the strength of the interaction between the system variable $`\widehat{X}`$ and the signal field. At the detector, the emitted signal can then be measured with a resolution limited by the quantum fluctuations of the field . If the signal is integrated over a finite time interval $`\mathrm{\Delta }t`$, this resolution is given by $$\frac{1}{\delta X^2}=4\gamma \mathrm{\Delta }t,$$ (2) where $`\delta X`$ is the expected error in the measurement result $`X_m`$ of the observable $`\widehat{X}`$ obtained during the time interval $`\mathrm{\Delta }t`$. The measurement result $`X_m`$ is now a classical record of the quantum property $`\widehat{X}`$. In quantum feedback, this record is used to condition the quantum dynamics of the system . In the case of linear feedback, the feedback can be described by a Hamiltonian of the form $$\widehat{H}_{\text{feedback}}(X_m)=2\mathrm{}\gamma X_m\widehat{Y}.$$ (3) If the time delay between the emission of the signal and the application of the feedback can be neglected, the effective dynamics of the system can be obtained by applying the operators representing the measurement and the feedback for the time interval $`\mathrm{\Delta }t`$ to both sides of the density matrix. The feedback dynamics of the density matrix can then be determined by averaging over all possible measurement results. Due to this averaging procedure, only quadratic terms in $`\widehat{X}`$ and $`\widehat{Y}`$ contribute, and the result reads $`{\displaystyle \frac{d}{dt}}\widehat{\rho }`$ $`=`$ $`\gamma \left(\widehat{X}\widehat{\rho }\widehat{X}{\displaystyle \frac{1}{2}}\widehat{X}^2\widehat{\rho }{\displaystyle \frac{1}{2}}\widehat{\rho }\widehat{X}^2\right)`$ (4) $`+`$ $`\gamma \left(\widehat{Y}\widehat{\rho }\widehat{Y}{\displaystyle \frac{1}{2}}\widehat{Y}^2\widehat{\rho }{\displaystyle \frac{1}{2}}\widehat{\rho }\widehat{Y}^2\right)`$ $`+`$ $`\gamma \left(i\widehat{Y}\widehat{\rho }\widehat{X}i\widehat{X}\widehat{\rho }\widehat{Y}+i\widehat{Y}\widehat{X}\widehat{\rho }i\widehat{\rho }\widehat{X}\widehat{Y}\right).`$ In this representation, the first term represents the original dephasing in $`\widehat{X}`$ caused by the emission of the signal, the second term represents the dephasing in $`\widehat{Y}`$ caused by the quantum noise in the feedback, and the third term represents the effects of the correlation between the measurement results obtained for $`\widehat{X}`$ and the feedback defined by $`\widehat{Y}`$. Note that the third term is not symmetric in $`\widehat{X}`$ and $`\widehat{Y}`$ due to the temporal sequence of the feedback. Where $`\widehat{X}`$ and $`\widehat{Y}`$ are applied to the same side of $`\widehat{\rho }`$, the measurement term $`\widehat{X}`$ is always applied before the feedback term $`\widehat{Y}`$. As will be shown below, this physically motivated sequence establishes a necessary correlation between the unitary evolution and the decoherence of the system. The dynamics given by equation (4) above can be transformed to a more conventional form by ordering the operator products according to the position of $`\widehat{\rho }`$, $`{\displaystyle \frac{d}{dt}}\widehat{\rho }`$ $`=`$ $`\gamma ((\widehat{X}+i\widehat{Y})\widehat{\rho }(\widehat{X}i\widehat{Y})`$ (5) $`{\displaystyle \frac{1}{2}}(\widehat{X}^2i2\widehat{Y}\widehat{X}+\widehat{Y}^2)\widehat{\rho }{\displaystyle \frac{1}{2}}\widehat{\rho }(\widehat{X}^2+i2\widehat{X}\widehat{Y}+\widehat{Y}^2)).`$ Here, the effects of measurement and feedback have been combined in such a way that their different physical origin is not recognizable anymore. In fact, the dynamics can now be summarized using only two operators, one to describe decoherence, and one to describe the unitary evolution of the quantum state. The dynamics then read $$\frac{d}{dt}\widehat{\rho }=\gamma \left(\widehat{c}\widehat{\rho }\widehat{c}^{}\frac{1}{2}\widehat{c}^{}\widehat{c}\widehat{\rho }\frac{1}{2}\widehat{\rho }\widehat{c}^{}\widehat{c}\right)\frac{i}{\mathrm{}}[\widehat{H}_{\text{eff.}},\widehat{\rho }],$$ (6) where the decoherence operator is $`\widehat{c}=\widehat{X}+i\widehat{Y}`$ and the effective Hamiltonian $`\widehat{H}_{\text{eff.}}`$ is given by $$\widehat{H}_{\text{eff.}}=\frac{\mathrm{}\gamma }{2}\left(\widehat{X}\widehat{Y}+\widehat{Y}\widehat{X}\right).$$ (7) It is interesting to compare this effective Hamiltonian with the actual feedback operator of equation (3). The essential difference is that the measurement value $`X_m`$ has now been replaced with the operator $`\widehat{X}`$. While this replacement appears to be a rather intuitive result since $`X_m`$ is the measurement result of the observable property represented by $`\widehat{X}`$, it is important to recognize that the replacement of a (generally continuous) real number with a hermitian operator completely changes the dynamics described by the Hamiltonian. In particular, the effective Hamiltonian is symmetric in $`\widehat{X}`$ and $`\widehat{Y}`$, indicating that the same Hamiltonian could be obtained from a measurement of $`\widehat{Y}`$ followed by a feedback in $`\widehat{X}`$. The difference between the two scenarios only appears in the exchanged roles of $`\widehat{c}`$ and $`\widehat{c}^{}`$. Equation (6) now gives the feedback dynamics of a system emitting a signal dependent on only a single observable of the system. The dynamics are therefore based on the case of an ideal quantum non-demolition measurement, as described by equation (1). Nevertheless the resulting decoherence operators $`\widehat{c}`$ in equation (6) have the form of annihilation operators, and the decoherence dynamics after feedback appears similar to typical photon emission dynamics. Interestingly, this observation has a simple classical explanation. In classical electrodynamics, radiation losses can be described by the back-action of the emitted electric field on the oscillating dipole itself. If the emitted dipole field is known, this kind of back-action can be represented by a feedback Hamiltonian. As reported in , homodyne detection of the emitted radiation can be used to identify and to compensate this back-action by applying a feedback Hamiltonian that cancels the effects of this deterministic back-action. It is thus possible to interpret the natural relaxation processes of a system interacting with the environment in terms of a hypothetical combination of minimal back-action measurement and quantum feedback, where the feedback represents the quantum version of classical back-action . The analysis given above shows how the classical information flow in a quantum feedback system can be identified with elements of the quantum dynamics of the density matrix. In the next section, we will consider the consequences of these results for interactions between two separate systems by analyzing the simulation of interaction Hamiltonians by quantum feedback. ## 3 Information dynamics in basic interactions In classical physics, interactions can always be described in terms of local forces acting on local systems. The fact that the forces originate from other systems can be taken into account by simply correlating the specific value of the force with the value of the corresponding system property. A classical interaction between a system $`A`$ and a system $`B`$ can thus be described as shown in figure 1: A force $`F_{AB}`$ depending on the value of the property $`X_A`$ of system $`A`$ acts locally on system $`B`$, while a force $`F_{BA}`$ depending on the value of the property $`X_B`$ of system $`B`$ acts locally on system $`A`$. The dynamics of each system can therefore be treated locally, while the connection between the systems is established by the transfer of the values of $`X_A`$ and of $`X_B`$ from one system to the other. This transfer of classical variables corresponds to a classical communication channel and could be realized by a classical feedback line using precise measurement data from the remote system. In the quantum case, things are a bit more complicated. As illustrated by figure 2, the Hamiltonian $`\widehat{H}_{AB}`$ describing the interaction dynamics necessarily acts on both systems at once. In general, it is not possible to define the local dynamics, as evidenced by the possibility of generating non-separable entangled states in the interaction. Effectively, the systems exchange genuine quantum information, and their interaction cannot be described in terms of local operations based on classical communication between the systems. As a result of this quantum communication, the density matrix $`\widehat{\rho }_{AB}`$ usually becomes entangled and cannot be represented by products of local density matrices. However, it is a well-known fact that entangled states become separable when a given amount of noise is added. Likewise, quantum interactions can be analyzed in terms of a feedback model closely corresponding to the classical case shown in figure 1 if the decoherence rate of the interacting systems is sufficiently high. Figure 3 shows the quantum feedback setup realizing an effective Hamiltonian of $`2\mathrm{}\gamma \widehat{X}_A\widehat{X}_B`$ by correlated local operations on systems $`A`$ and $`B`$. In this setup, the systems are now connected by classical communication lines carrying information on the measurement outcomes $`X_{m,A}`$ and $`\widehat{X}_{m,B}`$ for the system properties $`\widehat{X}_A`$ and $`\widehat{X}_B`$ from one system to the other. Using the quantum feedback theory introduced in section 2 above, it is then possible to determine the effective interaction dynamics between the systems. Specifically, the initial decoherence dynamics without feedback can be written as $`{\displaystyle \frac{d}{dt}}\widehat{\rho }_{AB}`$ $`=`$ $`D(\widehat{\rho }_{AB}),\text{with}`$ $`D(\widehat{\rho }_{AB})`$ $`=`$ $`\gamma \left(\widehat{X}_A\widehat{\rho }_{AB}\widehat{X}_A{\displaystyle \frac{1}{2}}\widehat{X}_A^2\widehat{\rho }_{AB}{\displaystyle \frac{1}{2}}\widehat{\rho }_{AB}\widehat{X}_A^2\right)`$ (8) $`+\gamma \left(\widehat{X}_B\widehat{\rho }_{AB}\widehat{X}_B{\displaystyle \frac{1}{2}}\widehat{X}_B^2\widehat{\rho }_{AB}{\displaystyle \frac{1}{2}}\widehat{\rho }_{AB}\widehat{X}_B^2\right).`$ where the indices of the operators indicate which system they act on. This decoherence permits a continuous measurement of $`\widehat{X}_A`$ and $`\widehat{X}_B`$ with a resolution of $`4\gamma \mathrm{\Delta }t`$ per time interval $`\mathrm{\Delta }t`$ as given by equation (2). The measurement results $`X_{m,A}`$ and $`X_{m.B}`$ obtained during each time interval can then be used to generate a linear feedback acting on the opposite system with feedback Hamiltonians given by $`\widehat{H}_{AB}(X_{m,A})`$ $`=`$ $`2\mathrm{}\gamma X_{m,A}\widehat{X}_B\text{and}`$ $`\widehat{H}_{BA}(X_{m,B})`$ $`=`$ $`2\mathrm{}\gamma X_{m,B}\widehat{X}_A.`$ (9) The feedback dynamics can then be determined most effectively by using equation (5), since a number of imaginary terms generated by the two feedback lines cancel. As a result, the decoherence term retains its original form, with the decoherence rate being doubled by the noise in the feedback. The joint dynamics of the feedback-coupled systems can then be written as $$\frac{d}{dt}\widehat{\rho }_{AB}=\frac{i}{\mathrm{}}[\widehat{H}_{\text{eff.}},\widehat{\rho }_{AB}]+2D(\widehat{\rho }_{AB}),$$ (10) where the effective interaction Hamiltonian is given by $$\widehat{H}_{\text{eff.}}=2\mathrm{}\gamma \widehat{X}_A\widehat{X}_B.$$ (11) The non-local quantum interaction represented by the Hamiltonian $`\widehat{H}_{\text{eff.}}`$ above can thus be simulated by a feedback setup in which only classical information is exchanged between the systems. The price to be paid for the replacement of quantum interactions with a classical signal transfer is given by the decoherence operator $`2D`$. The qualitative effects of the interaction Hamiltonian $`\widehat{H}_{\text{eff.}}`$ can then be analyzed in terms of an $`\widehat{X}_A`$-dependent unitary transform acting on system $`B`$ and a $`\widehat{X}_B`$-dependent unitary transform acting on $`A`$. More complicated interactions can be simulated if measurement information on other system variables is available in the emitted fields. In principle, any interaction Hamiltonian can be simulated by decomposing it into a sum of bilinear terms of the form $`2\mathrm{}\gamma \widehat{X}_A\widehat{X}_B`$, implementing each term by a separate feedback. Quantum feedback interactions can therefore be used to implement a wide range of interaction Hamiltonians between systems that are only connected by classical communication lines. ## 4 Interactions in a noisy environment: quantum limits of decoherence rates The special feature of an interaction realized entirely by quantum feedback is that the conditional evolution of the two systems is fully defined by the available classical information. In a direct quantum interaction between two systems, this kind of information is not necessarily available. However, it is possible that the interaction of the systems with the environment makes such information available even when the interaction is not implemented by feedback. In this case, the interaction is too noisy to entangle the system, and the classical information necessary to identify the local system dynamics is in principle available in the local environments of the interacting systems. Figure 4 illustrates this case: the measurement of $`X_A`$ in the emitted fields allows an identification of the Hamiltonian $`H_{AB}`$ determining the evolution of system $`B`$, and the measurement of $`X_B`$ in the emitted fields allows an identification of the Hamiltonian $`H_{BA}`$ determining the evolution of system $`A`$. The measurements in the environment can thus resolve the entanglement between the two systems and the environment by projecting the systems into a product state. As figure 4 suggests, the noise levels required to achieve this identification of local dynamics are equal to the noise levels generated by the corresponding quantum feedback based interaction. It is therefore possible to give some quantitative limits beyond which the information in the environment is definitely sufficient for an identification of the local dynamics. This feedback-based analysis can then be used to obtain quantitative results on the robustness of the entangling capabilities of interaction Hamiltonians against quantum noise . If the system dynamics is described by $$\frac{d}{dt}\widehat{\rho }_{AB}=\frac{i}{\mathrm{}}[\widehat{H}_0,\widehat{\rho }_{AB}]+L(\widehat{\rho }_{AB}),$$ (12) where $`\widehat{H}_0`$ is the Hamiltonian describing the unitary part of the dynamics and $`L`$ is the superoperator describing the non-unitary part of the dynamics, the condition for separability of the dynamics is given by a relation between the frequencies defining $`\widehat{H}_0/\mathrm{}`$ and the rates defining $`L`$. Using the results of section 3 above, it is possible to define the separability limit for the case of a bilinear interaction in the presence of dephasing between the eigenstates of the interaction Hamiltonian, $`\widehat{H}_0`$ $`=`$ $`\mathrm{}g_{AB}\widehat{X}_A\widehat{X}_B\text{and}`$ $`L(\widehat{\rho }_{AB})`$ $`=`$ $`\gamma _A\left(\widehat{X}_A\widehat{\rho }_{AB}\widehat{X}_A{\displaystyle \frac{1}{2}}\widehat{X}_A^2\widehat{\rho }_{AB}{\displaystyle \frac{1}{2}}\widehat{\rho }_{AB}\widehat{X}_A^2\right)`$ (13) $`+\gamma _B\left(\widehat{X}_B\widehat{\rho }_{AB}\widehat{X}_B{\displaystyle \frac{1}{2}}\widehat{X}_B^2\widehat{\rho }_{AB}{\displaystyle \frac{1}{2}}\widehat{\rho }_{AB}\widehat{X}_B^2\right),`$ where the strength of the interaction is defined by the coupling frequency $`g_{AB}`$ and the local decoherence rates of systems $`A`$ and $`B`$ are given by $`\gamma _A`$ and by $`\gamma _B`$, respectively. The separate variation of $`\gamma _A`$ and of $`\gamma _B`$ can be obtained by rescaling the operators $`\widehat{X}_A`$ and $`\widehat{X}_B`$ in equation (6) while leaving the product $`\widehat{X}_A\widehat{X}_B`$ unchanged. This equation of motion for the density matrix can then be separated into an effective local feedback scenario if (and only if) $$g_{AB}^2\gamma _A\gamma _B.$$ (14) That this condition is in fact both necessary and sufficient is a consequence of the conservation of eigenstates of $`\widehat{X}_A`$ and $`\widehat{X}_B`$ in the noisy dynamics given by equations (4). The only possible local feedback model consistent with this conservation of $`\widehat{X}_A`$ and $`\widehat{X}_B`$ is the model based on quantum non-demolition measurements of $`\widehat{X}_A`$ and $`\widehat{X}_B`$, as given by equations (3) to (11). Consequently, it is not possible to construct any local feedback scenario for equations (4) if $`g_{AB}^2>\gamma _A\gamma _B`$. In the case of a sum of several bilinear interactions in the Hamiltonian, sufficient conditions for separability can be obtained by simply combining all individual separability criteria. Obtaining a necessary conditions for separability is usually more difficult, since a general superoperator $`L`$ can have infinitely many possible decompositions . It is therefore generally unclear what selection of quantum measurements are optimal as a starting point for the feedback model. ## 5 Application to the two mode squeezing Hamiltonian It may now be instructive to consider a specific case of feedback induced interactions in optical systems. The most simple example is perhaps given by the dynamics of two resonant optical cavity modes described by the annihilation operators $`\widehat{a}_1=\widehat{x}_1+i\widehat{y}_1`$ and $`\widehat{a}_2=\widehat{x}_2+i\widehat{y}_2`$, respectively. If the attenuation rate of both cavities is $`\kappa `$, the initial dynamics of the two cavity modes is simply given by $$\frac{d}{dt}\widehat{\rho }_j=2\kappa \left(\widehat{a}_j\widehat{\rho }_j\widehat{a}_j^{}\frac{1}{2}\widehat{a}_j^{}\widehat{a}_j\widehat{\rho }_j\frac{1}{2}\widehat{\rho }_j\widehat{a}_j^{}\widehat{a}_j\right),$$ (15) describing the emission of light from the cavities. If the emitted light is detected by heterodyne detection, both quadrature components can be measured with a resolution of $$\frac{1}{\delta x^2}=\frac{1}{\delta y^2}=4\kappa \mathrm{\Delta }t.$$ (16) Therefore, both components can be used to generate feedback, where the cavity emission rate $`\kappa `$ corresponds to the decoherence rate $`\gamma `$ of sections 2 and 3. It is then possible to implement a two mode squeezing Hamiltonian by using the following feedback Hamiltonians, $`\widehat{H}_{x1x2}(x_{m,1})=+2\mathrm{}\kappa x_{m,1}\widehat{x}_2,`$ $`\widehat{H}_{x2x1}(x_{m,2})=+2\mathrm{}\kappa x_{m,2}\widehat{x}_1,`$ $`\widehat{H}_{y1y2}(y_{m,1})=2\mathrm{}\kappa y_{m,1}\widehat{y}_2,`$ $`\widehat{H}_{y2y1}(y_{m,2})=2\mathrm{}\kappa y_{m,2}\widehat{y}_1.`$ (17) If the total interaction dynamics of this feedback setup is written as $$\frac{d}{dt}\widehat{\rho }_{12}=\frac{i}{\mathrm{}}[\widehat{H}_{\text{eff.}},\widehat{\rho }_{12}]+2D(\widehat{\rho }_{12}),$$ (18) the effective two mode squeezing Hamiltonian is given by $$\widehat{H}_{\text{eff.}}=2\mathrm{}\kappa \left(\widehat{x}_1\widehat{x}_2\widehat{y}_1\widehat{y}_2\right)=\mathrm{}\kappa \left(\widehat{a}_1\widehat{a}_2+\widehat{a}_1^{}\widehat{a}_2^{}\right).$$ (19) In the absence of the decoherence represented by $`D`$, this two mode squeezing would squeeze the noise in the two mode quadratures $`\widehat{x}_1\widehat{x}_2`$ and $`\widehat{y}_1+\widehat{y}_2`$ to zero at an exponential relaxation rate of $`\kappa `$, creating standard squeezed state entanglement in the process. However, the decoherence effects of the feedback add noise to this relaxation process according to $`2D(\widehat{\rho }_{12})`$ $`=`$ $`3\kappa \left(\widehat{a}_1\widehat{\rho }_{12}\widehat{a}_1^{}{\displaystyle \frac{1}{2}}\widehat{a}_1^{}\widehat{a}_1\widehat{\rho }_{12}{\displaystyle \frac{1}{2}}\widehat{\rho }_{12}\widehat{a}_1^{}\widehat{a}_1\right)`$ (20) $`+\kappa \left(\widehat{a}_1^{}\widehat{\rho }_{12}\widehat{a}_1{\displaystyle \frac{1}{2}}\widehat{a}_1\widehat{a}_1^{}\widehat{\rho }_{12}{\displaystyle \frac{1}{2}}\widehat{\rho }_{12}\widehat{a}_1\widehat{a}_1^{}\right)`$ $`+3\kappa \left(\widehat{a}_2\widehat{\rho }_{12}\widehat{a}_2^{}{\displaystyle \frac{1}{2}}\widehat{a}_2^{}\widehat{a}_2\widehat{\rho }_{12}{\displaystyle \frac{1}{2}}\widehat{\rho }_{12}\widehat{a}_2^{}\widehat{a}_2\right)`$ $`+\kappa \left(\widehat{a}_2^{}\widehat{\rho }_{12}\widehat{a}_2{\displaystyle \frac{1}{2}}\widehat{a}_2\widehat{a}_2^{}\widehat{\rho }_{12}{\displaystyle \frac{1}{2}}\widehat{\rho }_{12}\widehat{a}_2\widehat{a}_2^{}\right).`$ This super operator still describes a relaxation of the field components at a rate of $`\kappa `$, but the feedback has doubled the rate at which quantum noise enters the cavities. The relaxation dynamics of the squeezed field fluctuations is therefore given by $`{\displaystyle \frac{d}{dt}}(\widehat{x}_1\widehat{x}_2)^2`$ $`=`$ $`4\kappa (\widehat{x}_1\widehat{x}_2)^2+2`$ $`{\displaystyle \frac{d}{dt}}(\widehat{y}_1+\widehat{y}_2)^2`$ $`=`$ $`4\kappa (\widehat{y}_1+\widehat{y}_2)^2+2,`$ (21) and the stationary solutions are exactly equal to the vacuum noise level of $`1/2`$. As expected, the feedback interaction therefore cannot entangle the cavity fields. However, it should be noted that even a slight increase in the squeezing interaction will reduce the noise level of both $`\widehat{x}_1\widehat{x}_2`$ and $`\widehat{y}_1+\widehat{y}_2`$ and lead to a violation of local uncertainties, indicating entanglement . The feedback induced interaction described above is therefore the strongest two mode squeezing interaction that can be realized by local operations and classical communication only. This result can be used to analyze the entangling capability of a two mode squeezing interaction in the presence of noise according to the procedure outlined in section 4 above. Specifically, a noisy two mode squeezing operation can be given in the form defined by equation (12). The dynamics is then described by the squeezing Hamiltonian $`\widehat{H}_0`$ and a decoherence operator $`L`$, $`\widehat{H}_0`$ $`=`$ $`2\mathrm{}g_{12}\left(\widehat{x}_1\widehat{x}_2\widehat{y}_1\widehat{y}_2\right)\text{and}`$ $`L(\widehat{\rho }_{12})`$ $`=`$ $`\gamma _{}\left(\widehat{a}_1\widehat{\rho }_{12}\widehat{a}_1^{}{\displaystyle \frac{1}{2}}\widehat{a}_1^{}\widehat{a}_1\widehat{\rho }_{12}{\displaystyle \frac{1}{2}}\widehat{\rho }_{12}\widehat{a}_1^{}\widehat{a}_1\right)`$ (22) $`+\gamma _+\left(\widehat{a}_1^{}\widehat{\rho }_{12}\widehat{a}_1{\displaystyle \frac{1}{2}}\widehat{a}_1\widehat{a}_1^{}\widehat{\rho }_{12}{\displaystyle \frac{1}{2}}\widehat{\rho }_{12}\widehat{a}_1\widehat{a}_1^{}\right)`$ $`+\gamma _{}\left(\widehat{a}_2\widehat{\rho }_{12}\widehat{a}_2^{}{\displaystyle \frac{1}{2}}\widehat{a}_2^{}\widehat{a}_2\widehat{\rho }_{12}{\displaystyle \frac{1}{2}}\widehat{\rho }_{12}\widehat{a}_2^{}\widehat{a}_2\right)`$ $`+\gamma _+\left(\widehat{a}_2^{}\widehat{\rho }_{12}\widehat{a}_2{\displaystyle \frac{1}{2}}\widehat{a}_2\widehat{a}_2^{}\widehat{\rho }_{12}{\displaystyle \frac{1}{2}}\widehat{\rho }_{12}\widehat{a}_2\widehat{a}_2^{}\right).`$ Here, the squeezing rate is expressed by the coupling frequency $`g_{12}`$ and the decoherence rates are expressed by $`\gamma _{}`$ for photon loss and by $`\gamma _+`$ for photon gain ($`\gamma _{}>\gamma _+`$ for stationary solutions). Comparison with equations (19) and (20) then shows that the two mode squeezing Hamiltonian could only originate from local measurements and feedback if $$g_{12}\gamma _+.$$ (23) This separability limit is consistent with the uncertainty limit obtained from the steady state of the squeezing dynamics, $`{\displaystyle \frac{d}{dt}}(\widehat{x}_1\widehat{x}_2)^2`$ $`=`$ $`(\gamma _{}\gamma _++2g_{12})(\widehat{x}_1\widehat{x}_2)^2+{\displaystyle \frac{1}{2}}(\gamma _{}+\gamma _+)`$ $`{\displaystyle \frac{d}{dt}}(\widehat{y}_1+\widehat{y}_2)^2`$ $`=`$ $`(\gamma _{}\gamma _++2g_{12})(\widehat{y}_1+\widehat{y}_2)^2+{\displaystyle \frac{1}{2}}(\gamma _{}+\gamma _+).`$ (24) Entanglement is obtained when the two quadrature components drop below the standard quantum limit, a result that is obtained in the steady state for $`g_{12}>\gamma _+`$. The example of two mode squeezing thus illustrates how quantum feedback scenarios can be used to identify the separability limit of a specific entanglement generating Hamiltonian. ## 6 Information inside and outside the system: separability and its interpretation The quantum feedback analysis of interactions between quantum systems illustrates the correspondence between classical interactions and quantum interactions by restoring the concept of local forces to the quantum formalism. Instead of treating the interaction process as an inseparable whole, it is now possible to analyze the sequence of cause and effect in the Hamiltonian dynamics. The difference between entanglement generating dynamics and separable dynamics is then a quantitative one, depending only on the noise levels of the systems. Nevertheless, there remains one significant difference that appears to introduce a qualitative element to the distinction between quantum systems and classical systems. In the completely classical interaction scenario shown in figure 1, we would naturally identify the properties of the systems directly with their measurement record in the environment. In the quantum case, however, this identification of measurement record and system property is not even legitimate in the separable case, since the resolution of quantum measurements is necessarily limited by quantum fluctuations. Consequently, it is not possible to access the “real” system properties at all, and a description of the systems in terms of the available information is all that we can achieve. In this sense, even local quantum systems do not have a classical description. What then is the significance of the distinction between separable and entangled systems? Quantum feedback provides some insight into this fundamental question by achieving a partial separation of information and physical causality. In the case of the separable interaction dynamics induced by quantum feedback, the information exchanged between the systems is available as classical information outside the systems. The information exchange between the systems is therefore completely detached from the actual system properties once the measurement has been performed. This detachment of information and physical properties is the decisive difference between the separable feedback and the direct entanglement generating interaction. In the pure Hamiltonian interaction, the exchange of quantum information implies that the information about the exact forces acting on each local system remains attached to the physical properties of the systems. No information about the interaction is available outside of the systems. Separability therefore indicates that the information about the interaction has detached itself from the quantum systems and is now available as classical information in the environment outside the systems, while entanglement generation indicates that the quantum information shared between the systems is exclusively confined inside the two interacting systems. ## 7 Conclusions As the analysis above has shown, the simulation of interaction Hamiltonians by quantum feedback can provide fundamental insights into the dynamics of information exchange between interacting systems. In particular, the dynamics expressed by an interaction Hamiltonian can be interpreted in terms of a completely classical information exchange if the decoherence rates are comparable to the frequencies defining the strength of the interaction in the Hamiltonian. The difference between a separable interaction that can be represented by local operations and classical communication between the systems and a genuine entanglement generating quantum interaction can then be expressed quantitatively in terms of the decoherence rates associated with the availability of information in the environment outside the interacting systems. It may thus be possible to identify the robustness of quantum interactions against various noise effects by applying an appropriate quantum feedback model. Part of this work has been supported by the JST-CREST project on quantum information processing.
warning/0506/cond-mat0506610.html
ar5iv
text
# Spin properties of single electron states in coupled quantum dots ## I Introduction The possibility of tuning spin-orbit coupling Datta and Das (1990); Nitta et al. (1997); Koga et al. (2002) in low-dimensional semiconductor electronic structures has stirred great interest in spin properties of lateral semiconductor electron systems in the presence of Dresselhaus Dresselhaus (1955) and Bychkov-Rashba Bychkov and Rashba (1984); Rashba (1960) spin-orbit couplings. The former appears in low-dimensional systems lacking inversion symmetry in the bulk (such as zinc-blende semiconductors), the latter in low-dimensional structures with asymmetric confining potentials. The principal question is what spin and charge properties and to what degree can be affected and manipulated by this tuning. Such questions are of fundamental importance for spintronics.Žutić et al. (2004) Electron spins in coupled quantum dot systems have been proposed to perform universal gating of quantum computers. Loss and DiVincenzo (1998) The spin acts as a qubit and exchange coupling provides the physical realization of two-qubit gates. Burkard et al. (1999); Hu and Das Sarma (2000) Another application of a controlled coupling between dots is spin entanglement distillation in which singlet and triplet states get spatially separated during adiabatic passage through trapped states. Fabian and Hohenester (2004) Understanding of spin-orbit properties of coupled dots is thus of great interest to quantum information processing. Spin-orbit coupling provides a way for orbital degrees of freedom to influence spin states. As a result the spin dynamics is affected, making spin qubit operations more complex (it was shown, though, that two qubit operations can be performed reliably even in the presence of spin-orbit interaction which leads to anisotropic exchangeBurkard and Loss (2002); Stepanenko et al. (2003)). Furthermore, spin-orbit coupling leads to spin decoherence and relaxation due to phonons, Khaetskii and Nazarov (2001); Destefani and Ulloa (2004); Woods et al. (2002); Cheng et al. (2004); Golovach et al. (2004); Bulaev and Loss (2005) limiting the operation time. The impressive experimental progress in coherent oscillations in coupled dot systems Petta et al. (2004a); Hayashi et al. (2003); van der Wiel et al. (2003); Petta et al. (2004b), as well in spin dephasing and spin manipulation in single Hanson et al. (2003, 2005) and double dots Johnson et al. (2005), provides additional strong impetus for investigating spin states in double dots. Spin-orbit properties of single dots have been already extensively investigated. Voskoboynikov et al. (2001); Governale (2002); de Sousa and Sarma (2003); Valin-Rodriguez et al. (2004); Destefani et al. (2004a, b); Kuan et al. (2004); Tsitsishvili et al. (2004); Voskoboynikov et al. (2003); Bulgakov and Sadreev (2001); Pfeffer and Zawadzki (1999); Destefani and Ulloa (2005) In this paper we investigate the role of spin-orbit coupling, represented by the Dresselhaus (both linear and cubic) and Bychkov-Rashba terms, in spin and charge properties of two laterally coupled quantum dots based on GaAs materials parameters. We perform numerically exact calculations of the energy spectrum using the method of finite differences. We first study the general structure of the energy spectrum and the spin character of the states of the double dot system. We construct the group theoretical correlation diagram for the single and double dot states and indicate the possible transitions due to spin-orbit coupling. This group theoretical classification is used in combination with Löwdin perturbation theory to explain analytically our numerical results. In particular, we show that while allowed by symmetry, the specific forms of the linear spin-orbit interactions do not lead to spin hot spots in the absence of magnetic field (spin hot spots are nominally degenerate states lifted by spin-orbit couplingFabian and S. Das Sarma (1998)). Only the cubic Dresselhaus term gives spin hot spots. If identified experimentally, the strength of the cubic term can be detected. We next focus on two important measurable parameters: electronic g-factor and tunneling amplitude. In single dots the variation of the effective g-factor with the strength of the spin-orbit interaction has been investigated earlier.de Sousa and Sarma (2003) The effect is not large, amounting to a fraction of a percent. Similar behavior is found for double dots. In our case of GaAs the contribution to the g-factor from spin-orbit coupling is typically about 1%, due to the linear Dresselhaus term. More exciting is the prospect of influencing coherent tunneling oscillations between the dots by modulating the spin-orbit coupling strength. Two effects can appear: (i) the tunneling amplitude or frequency can be modulated by spin-orbit coupling and, (ii) the tunneling amplitude can be spin dependent. We show how a naive application of perturbation theory leads to a misleading result that (i) is present in the second order in linear spin-orbit coupling strengths, giving rise to an effective tunneling Hamiltonian involving spin-flip tunneling at zero magnetic field. Both numerical calculations and an analytical argument, presented here, show that this is incorrect and that there is no correction to the tunneling Hamiltonian in the second order of linear spin-orbit terms. The dominant correction in the second order comes from the interference of linear and cubic Dresselhaus terms. We propose to use this criterion, that the corrections to linear terms vanish in the second order, to distinguish between single and double dots as far as spin properties of the states are concerned. Indeed, at very small and very large intradot couplings the states have a single dot character and the correction to energy due to linear spin orbit terms depends on the interdot distance (except for the two lowest states which provide tunneling). We find that dots are “coupled” up to the interdot distance of about five single-dot confinement lengths. In the presence of magnetic field the time reversal symmetry is broken. The presence of spin-orbit coupling then in general leads to a spin dependent tunneling amplitude. Spin up and spin down states will oscillate between the two dots with different frequencies (for our GaAs dots the relative difference of the frequencies is at the order of 0.1%, but is higher in materials with larger spin-orbit coupling). This leads to a curious physical effect, namely, that of a spatial separation of different spin species. Indeed, starting with an electron localized on one dot, with a spin polarized in the plane (that is, a superposition of up and down spins), after a sequence of coherent oscillations the electron state is a superposition of spin up localized on one, and spin down localized on the other dot. A single charge measurement on one dot collapses the wave function to the corresponding spin state, realizing a spin to charge conversion. We construct an effective, four state (two spin and two sites) tunneling Hamiltonian for the single electron double dot system, which takes into effect the above results. Such a Hamiltonian should be useful for constructing realistic model theories of spin dephasing, spin tunneling, and kinetic exchange coupling in coupled quantum dot systems. The paper is organized as follows. In Sec. II we introduce the model, the numerical technique, and materials and system parameters. In Sec. III we review the benchmark case of single dots with spin-orbit coupling and magnetic field. Coupled double dots are studied in Sec. IV, separately in zero and finite magnetic fields. We conclude with the discussion of our results in Sec. V. ## II Model We consider conduction electrons confined in a plane of a zinc-blende semiconductor heterostructure, with additional confinement into lateral dots given by appropriately shaped top gates. A magnetic field B is applied perpendicular to the plane. In the effective mass approximation the single-electron Hamiltonian of such a system, taking into account spin-orbit coupling, can be decomposed into several terms: $$H=T+V_C+H_Z+H_{BR}+H_D+H_{D3}.$$ (1) Here $`T=\mathrm{}^2\text{K}^2/2m`$ is the kinetic energy with the effective electron mass $`m`$ and kinetic momentum $`\mathrm{}\text{K}=\mathrm{}\text{k}+e\text{A}=i\mathrm{}+e\text{A}`$; $`e`$ is the proton charge and $`\text{A}=B(y/2,x/2,0)`$ is the vector potential of $`\text{B}=(0,0,B)`$. Operators of angular momentum with mechanical and canonical momenta are denoted as $`\text{L}=\text{r}\times (\mathrm{}\text{K})`$ and $`\text{l}=\text{r}\times (\mathrm{}\text{k})`$. The quantum dot geometry is described by the confining potential $`V_C(\text{r})`$. Single dots will be described here by a parabolic potential $`V_C=(1/2)m\omega _0^2r^2`$, characterized by confinement energy $`E_0=\mathrm{}\omega _0/2`$ and confinement length $`l_0=(\mathrm{}/m\omega _0)^{1/2}`$, setting the energy and length scales, respectively. Coupled double dots will be described by two displaced (along $`𝐱`$) parabolas: $$V_C^{dd}=\frac{1}{2}m\omega _0^2[(|x|l_0d)^2+y^2];$$ (2) the distance between the minima is $`2d`$ measured in the units of $`l_0`$. The Zeeman energy is given by $`H_Z=(g^{}/2)\mu _B\sigma _zB`$, where $`g^{}`$ is the conduction band g-factor, $`\mu _B`$ is the Bohr magneton, and $`\sigma _z`$ is the Pauli matrix. In order to relate the magnetic moment of electrons to their orbital momentum, we will use dimensionless parameter $`\alpha _Z=g^{}m/2m_e`$, where $`m_e`$ is the free electron mass. Spin-orbit coupling gives additional terms in confined systems.Žutić et al. (2004) The Bychkov-Rashba Hamiltonian,Rashba (1960); Bychkov and Rashba (1984) $$H_{BR}=\stackrel{~}{\alpha }_{BR}\left(\sigma _xK_y\sigma _yK_x\right),$$ (3) appears if the confinement is not symmetric in the growth direction (here z). The strength $`\stackrel{~}{\alpha }_{BR}`$ of the interaction can be tuned by modulating the asymmetry by top gates. Due to the lack of spatial inversion symmetry in zinc-blende semiconductors, the spin-orbit interaction of conduction electrons takes the form of the Dresselhaus Hamiltonian Dresselhaus (1955) which, when quantized in the growth direction z of our heterostructure gives two terms, one linear and one cubic in momentum:Dyakonov and Kachorovskii (1986) $`H_D`$ $`=`$ $`\gamma _cK_z^2\left(\sigma _xK_x+\sigma _yK_y\right),`$ (4) $`H_{D3}`$ $`=`$ $`(\gamma _c/2)\left(\sigma _xK_xK_y^2\sigma _yK_yK_x^2\right)+H.c.,`$ (5) where $`\gamma _c`$ is a material parameter. The angular brackets in $`H_D`$ denote quantum averaging in the z direction–the magnitude of $`H_D`$ depends on the confinement strength. We will denote the sum of the two linear spin-orbit terms by $`H_{\mathrm{lin}}=H_D+H_{BR}`$. The complete spin-orbit coupling is then $`H_{SO}=H_{\mathrm{lin}}+H_{D3}`$. We find it useful to introduce dimensionless strengths of the individual terms of the spin-orbit interaction by relating them to the confinement energy of a single dot $`E_0`$. We denote $`\alpha _{BR}=\stackrel{~}{\alpha }_{BR}/E_0l_0`$ and $`\alpha _D=\gamma _ck_z^2/E_0l_0`$ for linear terms, and $`\alpha _{D3}=\gamma _c/2E_0l_0^3`$ for the cubic Dresselhaus term. In our numerical examples we use $`E_0=\mathrm{}\omega _0/2=1.43`$ meV for the confinement energy, which corresponds to the confinement length of $`l_0=20`$ nm. We further use bulk GaAs materials parameters: $`m=0.067m_e`$, $`g^{}=0.44`$, and $`\gamma _c=27.5`$ eV$`\mathrm{\AA }^3`$. For $`k_z^2`$ we choose $`5.3\times 10^4\mathrm{\AA }^2`$, which corresponds to $`\gamma _ck_z^2=14.6`$ meV Å. This value of $`k_z^2`$ corresponds to the ground state of a 6 nm thick triangular potential well.de Sousa and Sarma (2003) For $`\stackrel{~}{\alpha }_{BR}`$ we choose a generic value of 4.4 meVÅ, which is in line of experimental observations.Miller et al. (2003); Knap et al. (1996) The dimensionless parameter of the Zeeman splitting then is $`\alpha _Z=0.015`$, while the relative strengths of the spin-orbit interactions are $`\alpha _{BR}0.015`$, $`\alpha _D0.05`$, and $`\alpha _{D3}0.001`$. Except for anti-crossings, the spin-orbit interaction is a small perturbation to the electronic structure; it is, however, essential for investigating spin properties. Our analytical calculations will often refer to the Fock-Darwin Fock (1928); Darwin (1931) spectrum, which is the spectrum of Hamiltonian (1) for a single dot with $`H_{SO}=0`$. The corresponding wave functions $`\mathrm{\Psi }`$ (expressed in polar coordinates $`r`$ and $`\varphi `$), and energies $`ϵ`$ are $`\mathrm{\Psi }_{n,l,\sigma }(r,\varphi )`$ $`=`$ $`C\rho ^{|l|}e^{\rho ^2/2}L_n^{|l|}(\rho ^2)e^{il\varphi }\xi (\sigma ),`$ (6) $`ϵ_{n,l,\sigma }`$ $`=`$ $`2E_0{\displaystyle \frac{l_0^2}{l_B^2}}(2n+|l|+1)+B{\displaystyle \frac{\mathrm{}e}{2m}}(l+\alpha _Z\sigma ),`$ (7) where $`\rho =r/l_B`$ is the radius in the units of the effective confinement length $`l_B`$, defined by $`l_B^2=l_0^2/\sqrt{(1+B^2e^2l_0^4/4\mathrm{}^2)}`$; $`n`$ and $`l`$ are the principal and orbital quantum numbers, respectively, $`C`$ is the state dependent normalization constant, and $`L_n^{|l|}`$ are associated Laguerre polynomials. Spinors $`\xi (\sigma )`$ describe the spin $`\sigma `$ state of the electrons. Since the parabolic dot has rotational symmetry in the plane, the states have well defined orbital momentum $`l`$ and spin $`\sigma `$ in the z-direction. We also introduce a useful dimensionless measure $`\theta `$ of the strength of the magnetic field induced confinement compared to the potential confinement: $`\theta =Bel_B^2/2\mathrm{}`$, $`0<\theta <1`$. The parameter $`\theta `$ gives the number of magnetic flux quanta through a circle with radius $`l_B`$. For large magnetic fields $`\theta 1(Bel_0^2/2\mathrm{})^2/2`$. The confining length can be expressed as $`l_B=l_0(1\theta ^2)^{1/4}`$. As it is not possible to solve for the spectrum of Hamiltonian (1) analytically, we treat it numerically with the finite differences method using Dirichlet boundary conditions (vanishing of the wave function at boundaries). The magnetic field is included via the Peierls phase: if $`H(\text{r}_i,\text{r}_j)`$ is the discretized Hamiltonian connecting grid points $`\text{r}_i`$ and $`\text{r}_j`$ at $`B=0`$, the effects of the field are obtained by adding a gauge phase: $`H(\text{r}_i,\text{r}_j)\mathrm{exp}[i(e/\mathrm{})_{\text{r}_i}^{\text{r}_j}\text{A}.\mathrm{d}\text{l}]`$. In our simulations we typically use $`50\times 50`$ grid points. The resulting matrix eigenvalue problem is solved by Lanczos diagonalization. The achieved accuracy is about $`10^5`$. ## III Single dots As a starting point we review the effects of spin-orbit coupling in single dots. We are interested in the changes to the spectrum and, in particular, to the magnetic moment of the lowest states, that is, to the effective $`g`$-factor. The calculated spectrum of a single dot is shown in Fig. 1. There are three ways in which spin-orbit coupling influences the spectrum: (i) First, the levels are uniformly shifted down, in proportion to $`\alpha ^2`$ (by $`\alpha `$ here we mean any of $`\alpha _{BR}`$, $`\alpha _D`$, or $`\alpha _{D3}`$). (ii) Second, the degeneracy at $`B=0`$ is lifted, again in proportion to $`\alpha ^2`$ (1b). (iii) Finally, at some magnetic field the level crossing of the Fock-Darwin levels is lifted by spin-orbit coupling. The resulting level repulsion is linear in $`\alpha `$ (1c). The states here are the spin hot spots, that is states in which both Pauli spin up and down species contribute significantly.Fabian and S. Das Sarma (1998); Bulaev and Loss (2005); Destefani et al. (2004b) The above picture can be understood from general symmetry considerations within the framework of perturbation theory. All spin-orbit terms commute, at $`B=0`$, with the time inversion operator $`T=i\sigma _y\widehat{C}`$, where $`\widehat{C}`$ is the operator of complex conjugation. Therefore Kramer’s degeneracy is preserved so that the states are always doubly degenerate. The linear terms can be transformed into each other by a unitary transformation $`(\sigma _x+\sigma _y)/\sqrt{2}`$ (spin rotation around $`[110]`$ by $`\pi `$ ), which commutes with $`H_0`$. Therefore the effects on the energy spectrum induced individually by the linear Dresselhaus and the Bychkov-Rashba terms are identical at $`B=0`$. At finite magnetic fields the two interactions give qualitatively different effects in the spectrum, especially for spin hot spots, as discussed below. For any $`B`$ the following commutation relations hold for the linear terms: $$[H_{BR},l_z+s_z]=0,[H_D,l_zs_z]=0.$$ (8) This means that $`H_{BR}`$ commutes with the angular momentum, while $`H_D`$ does not. This will influence the interference between the two terms in the energy spectrum. We can use the Fock-Darwin states as a basis for perturbation theory. Up to the second order the energy of state $`|i=\mathrm{\Psi }_{n,l,\sigma }`$ is $$E_i=ϵ_i+i|H_{SO}|i+\underset{ji}{}\frac{i|H_{SO}|jj|H_{SO}|i}{ϵ_iϵ_j}.$$ (9) The first order correction is zero for all spin-orbit terms since $`H_{SO}`$ contain only odd powers of K whose expectation values in the Fock-Darwin states vanish. If the perturbation expansion is appropriate, the spin-orbit interactions have a second order (in $`\alpha `$) effect on energy. Both linear spin-orbit terms couple states with orbital momenta $`l`$ differing by one. It then follows from the commutation relations (8) that $`H_{BR}`$ preserves the total angular momentum $`l+s`$, while $`H_D`$ preserves the quantity $`ls`$. As a result, there is no correction to the energy from the interference terms between $`H_{BR}`$ and $`H_D`$ in Eq. (9): $`i|H_{BR}|jj|H_D|i=0`$. As for the cubic Dresselhaus term, only the following orbital states are coupled: $`(l,)\{(l+3,),(l1,)\}`$ and $`(l,)\{(l3,),(l+1,)\}`$. Due to these selection rules there are no interference terms $`H_{D3}H_{BR}`$, but terms $`H_{D3}H_D`$ will contribute to energy perturbation. The Bychkov-Rashba and Dresselhaus Hamiltonians act independently on the Fock-Darwin spectrum (up to the second order). To gain more insight into the perturbed structure of the spectrum at $`B=0`$, we rewrite Eq. (9) using an auxiliary anti-hermitian operator $`H_{SO}^{op}`$ defined by the commutation relation $`[H_0,H_{SO}^{op}]=H_{SO}`$. If such an operator exists, the second order correction in (9) is then $`{\displaystyle \underset{j𝒩}{}}{\displaystyle \frac{i|H_{SO}|jj|H_{SO}|i}{ϵ_iϵ_j}}=i|{\displaystyle \frac{1}{2}}[H_{SO}^{op},H_{SO}]|i+`$ $`+Re(i|H_{SO}P_𝒩H_{SO}^{op}|i),`$ (10) where $`P_𝒩`$ is the projector on the subspace $`𝒩`$ of the states excluded from the summation. In our case here it is just one state, $`𝒩=\{|i\}`$. The last term in (10) then vanishes. The auxiliary operator for $`H_{D3}`$ is not known and if found, it must depend on the confining potential. Operators for the linear terms are:Aleiner and Faľko (2001) $`H_D^{op}`$ $`=`$ $`i(\alpha _D/2l_0)(x\sigma _xy\sigma _y),`$ (11) $`H_{BR}^{op}`$ $`=`$ $`i(\alpha _{BR}/2l_0)(y\sigma _xx\sigma _y).`$ (12) The corresponding commutators are (in the zero magnetic field $`\text{K}=\text{k},L_z=l_z,\theta =0`$; the last expression will be useful later) $`[H_D^{op},H_D]`$ $`=`$ $`E_0\alpha _D^2(1\sigma _zL_z),`$ (13) $`[H_{BR}^{op},H_{BR}]`$ $`=`$ $`E_0\alpha _{BR}^2(1+\sigma _zL_z),`$ (14) $`[H_D^{op},H_{D3}]`$ $`=`$ $`E_0l_0^2\alpha _D\alpha _{D3}(K_x^2+K_y^2`$ (15) $`2\sigma _z[xK_yK_x^2yK_xK_y^22i\theta (xK_x+yK_y)]).`$ Because $`[H_D^{op},H_{BR}]+[H_{BR}^{op},H_D]=0`$, the corrections to the second order perturbation add independently for $`H_{BR}`$ and $`H_D`$ (as also noted above from the selection rules), we can introduce $`H_{\mathrm{lin}}^{op}=H_D^{op}+H_{BR}^{op}`$. An alternative route to Eq. (10) is to transform the Hamiltonian withAleiner and Faľko (2001) $`U=\mathrm{exp}(H_{SO}^{op})`$ to $`\stackrel{~}{H}=H_0(1/2)[H_{SO},H_{SO}^{op}]`$ in the second order of $`\alpha `$. The final result can be also obtained in a straightforward way by using the Thomas-Reiche-Kuhn sum rule in the second order of perturbation theory with the original spin-orbit terms. The resulting effective Hamiltonian is (terms depending on $`\alpha _{D3}`$ are omitted here) $$\stackrel{~}{H}=H_0E_0(\alpha _D^2+\alpha _{BR}^2)/2+E_0\sigma _zL_z(\alpha _D^2\alpha _{BR}^2)/2.$$ (16) This Hamiltonian, in which the spin-orbit coupling appears in its standard form, neatly explains point (ii) about the lifting of the degeneracy at $`B=0`$. The levels in Fig. 1b, for example, are four fold degenerate ($`|l|=1`$, $`|\sigma |=1`$) without spin-orbit coupling. Turning on, say, $`H_D`$, will split the levels into two groups: energy of the states with $`l\sigma >0`$ would not change in the second order, while the states with $`l\sigma <0`$ will go down in energy by $`E_0\alpha _D^2`$, as seen in Fig. 1b. ### III.1 Spin hot spots Spin hot spots are states formed by two or more states whose energies in the absence of spin-orbit coupling are degenerate or close to being degenerate, while turning on the coupling removes the degeneracy.Fabian and S. Das Sarma (1998) Such states are of great importance for spin relaxation, which is strongly enhanced by their presence.Fabian and S. Das Sarma (1999); Bulaev and Loss (2005) The reason is that the degeneracy lifting mixes spin up and spin down states and so transitions between states of opposite magnetic moment will involve spin flips with a much enhanced probability compared to normal states. Figure 1c shows an interesting situation where two degenerate levels are lifted by spin-orbit coupling.Destefani et al. (2004b); Bulaev and Loss (2005) The lifting is of the first order in $`\alpha `$, unlike the lifting of degeneracy at $`B=0`$ in which case the degenerate states are not directly coupled by $`H_{SO}`$. In a finite magnetic field, at a certain value $`B_{\mathrm{acr}}`$, the states of opposite spins and orbital momenta differing by one cross each other, as follows from the equation (7). The crossing field is $`B_{\mathrm{acr}}|\alpha _Z|^{1/2}\mathrm{}/(el_0^2)`$, which is about 13.4 T for our parameters (making the confinement length larger the magnitude of the field would decrease). Spin-orbit interaction couples the two states thereby lifting the degeneracy. For GaAs, where $`g^{}<0`$, only the Bychkov-Rashba term couples the two states. The Dresselhaus terms are not effective ($`H_{D3}`$ would introduce such a splitting at $`3B_{\mathrm{acr}}`$). The energy splitting due to $`H_{BR}`$ is $$\mathrm{\Delta }=cE_0\alpha _{BR}(1\theta _{\mathrm{acr}})(1\theta _{\mathrm{acr}}^2)^{1/4},$$ (17) where $`c`$, which is a number of order 1, depends on the quantum numbers of the two states. Since spin hot spots at $`B_{\mathrm{acr}}`$ are due only $`H_{BR}`$, the splittings could help to sort out the Bychkov-Rashba versus Dresselhaus contributions. Figure 1c shows the calculated level repulsion for states $`n=0,l=0,\sigma =`$ and $`n=0,l=1,\sigma =`$. The magnitude of $`\mathrm{\Delta }`$, though being linear in $`\alpha _{BR}`$, is on the order of $`10^3`$ meV and thus comparable to the energy scales associated with quadratic spin-orbit perturbations. ### III.2 Effective g-factor When probing spin states in quantum dots with magnetic field, important information comes from the measured Zeeman splitting. We will focus here on the two lowest spin states and calculate the effective $`g`$-factor as $`g=(E_{0,0,}E_{0,0,})/(\mu _BB)`$. If $`H_{SO}=0`$, then the effective $`g`$-factor equals to the conduction band value $`g^{}`$. The actual value in the presence of spin-orbit coupling is important for understanding single spin precession in magnetic field, which seems necessary to perform single qubit operations in quantum dots. We have obtained the following contributions to the $`g`$-factor from non-degenerate (that is, excluding spin hot spots) second-order perturbation theory \[Eq. (9)\] (for linear spin-orbit terms these are derived also inde Sousa and Sarma (2003); Valín-Rodriguez et al. (2002)): $`\delta g_{DD}`$ $`=`$ $`{\displaystyle \frac{\alpha _D^2}{2m/m_e}}{\displaystyle \frac{\sqrt{1\theta ^2}[1\theta ^22(1+\theta ^2)\alpha _Z]}{1\theta ^2(1+4\alpha _Z+4\alpha _Z^2)}},`$ $`\delta g_{BRBR}`$ $`=`$ $`{\displaystyle \frac{\alpha _{BR}^2}{2m/m_e}}{\displaystyle \frac{\sqrt{1\theta ^2}[1\theta ^2+2(1+\theta ^2)\alpha _Z]}{1\theta ^2(14\alpha _Z+4\alpha _Z^2)}},`$ $`\delta g_{DD3}`$ $`=`$ $`{\displaystyle \frac{\alpha _D\alpha _{D3}}{m/m_e}}{\displaystyle \frac{(1+\theta ^2)[1\theta ^22(1+\theta ^2)\alpha _Z]}{1\theta ^2(1+4\alpha _Z+4\alpha _Z^2)}},`$ $`\delta g_{D3D3}`$ $`=`$ $`{\displaystyle \frac{\alpha _{D3}^2}{8m/m_e\theta \sqrt{1\theta ^2}}}({\displaystyle \frac{2(1\theta )^2(1+\theta ^2)^2}{1\theta (1+2\alpha _Z)}}+`$ (18) $`+`$ $`{\displaystyle \frac{(1\theta )^4(1+\theta )^2}{3\theta (1+2\alpha _Z)}}+{\displaystyle \frac{3(1\theta )^6}{3\theta (32\alpha _Z)}}+`$ $`+`$ $`{\displaystyle \frac{3(1+\theta )^6}{3+\theta (32\alpha _Z)}}{\displaystyle \frac{2(1+\theta )^2(1+\theta ^2)^2}{1+\theta (1+2\alpha _Z)}}`$ $``$ $`{\displaystyle \frac{(1\theta )^2(1+\theta )^4}{3+\theta (1+2\alpha _Z)}}).`$ Here $`\delta g_{AB}`$ stands for a correction that is proportional to $`\alpha _A\alpha _B`$. The functions (18) are plotted in Fig. 2. We can understand the limits of $`\delta g`$ at $`B\mathrm{}(\theta 1)`$ if we notice that in the natural length unit $`l_B`$ the momentum $`K_x=i_xyBe/2\mathrm{}=l_B^1[i_{x/l_B}\theta (y/l_B)]`$. In the limit $`B\mathrm{}`$ the matrix elements of $`H_D`$, which is linear in $`K`$, scale as $`l_B^1`$, while the Fock-Darwin energies scale as $`l_B^2`$. The second order $`D`$-$`D`$ correction to $`E_{0,0,}E_{0,0,}`$ is thus independent of $`l_B`$; it converges to $`E_0\alpha _D^2/(1+\alpha _Z)`$. The $`BR`$-$`BR`$ correction is analogous, with the limit $`E_0\alpha _{BR}^2/(1\alpha _Z)`$. To get the g-factor we divide the energy differences by $`\mu _BB`$ and get $`\delta g_{DD}(\theta 1)B^1`$; similarly for $`H_{BR}`$. Since $`H_{D3}`$ scales as $`l_B^3`$ one gets $`\delta g_{DD3}(\theta 1)2\alpha _D\alpha _{D3}m/(1+\alpha _Z)m_e`$ and $`\delta g_{D3D3}(\theta 1)B`$. It seems that for increasing $`B`$ there inevitably comes a point where the influence of $`H_{D3}`$ on the g-factor dominates. But at $`B=B_{\mathrm{acr}}`$ there is an anti-crossing of the states $`(0,0,)`$ and $`(0,1,)`$ so for larger $`B`$ the g-factor does not describe the energy difference between the two lowest states, but between the second excited state and the ground state. The value of $`B`$ where $`\delta g_{D3D3}=\delta g_{DD}`$ is given by $`B(\mathrm{}/el_0^2)\alpha _D/\alpha _{D3}\sqrt{2}`$. For GaAs parameters it is $`25`$ T. After anti-crossing the first exited state is $`\mathrm{\Psi }_{0,1,}`$ and its energy difference to the ground state is divided by $`\mu _BB`$ is $`1/B^2`$ for $`H_0`$. The corrections from spin-orbit terms are $`D`$-$`D,BR`$-$`BR1/B^5`$, $`D`$-$`D31/B^4`$, and $`D3`$-$`D31/B^3`$. ## IV Double dots A double dot structure comprises two single dots close enough for their mutual interaction to play an important role. Here we consider symmetric dots modeled by $`V_C^{dd}`$ of Eq. (2). Such a potential has an advantage that in the limits of small $`d0`$ and large $`d\mathrm{}`$, the solutions converge to the single dot solutions centered at $`x=0`$ and $`\pm l_0d`$, respectively. We denote the displaced Fock-Darwin states as $`\mathrm{\Psi }_{n,l,\sigma }^{\pm d}(x,y)\mathrm{\Psi }_{n,l,\sigma }(x\pm l_0d,y)`$. The symmetries of the double dot Hamiltonian with spin-orbit couplings are listed in Tab. 1. The time symmetry is always present at $`B=0`$, giving Kramer’s double degeneracy. The rotational space symmetry from the single dot case is lost; instead there are two discrete symmetries – reflections $`I_x`$ about $`y`$ and $`I_y`$ about $`x`$. In zero magnetic field and without spin-orbit terms, the Hamiltonian has both $`I_x`$ and $`I_y`$ symmetries. If only Rashba or Dresselhaus terms are present, we can still preserve symmetries $`I_x`$ and $`I_y`$ by properly defining the symmetry operators to act also on the spinors (forming the double group). The Bychkov-Rashba term, $`H_0+H_{BR}`$, is invariant to operations defined by the spatial invariance. This is not the case of $`H_D`$, since here the operators $`i\sigma _yI_x`$ and $`i\sigma _xI_y`$ do not describe a spatial reflection of both the orbital and spinor parts. The symmetry operations for $`H_{BR}`$ and $`H_D`$ are connected by the unitary transformation $`(\sigma _x+\sigma _y)/\sqrt{2}`$, which connects the two Hamiltonians themselves. Finally, if both spin-orbit terms are present, or at $`B>0`$, the only space symmetry left is $`I=I_xI_y`$. In the following we consider separately the cases of zero and finite magnetic fields. ### IV.1 Energy spectrum in zero magnetic field, without spin-orbit terms. If no spin-orbit terms are present the group of our double dot Hamiltonian is $`C_{2v}SU(2)`$. The $`SU(2)`$ part accounts for the (double) spin degeneracy. The orbital parts of the eigenstates of the Hamiltonian therefore transform according to the irreducible representations of $`C_{2v}`$. The representationsKoster (1957) $`\mathrm{\Gamma }_i`$, $`i=1\mathrm{}4`$, along with their transformation properties under the symmetries of $`C_{2v}`$, are listed in Tab. 2. The symmetry properties will be used in discussing the perturbed spectrum. We denote the exact eigenfunctions of the double dot Hamiltonian with $`H_{SO}=0`$ as $`\mathrm{\Gamma }_{i\sigma }^{ab}`$ where $`a(b)`$ is the single dot level to which this eigenfunction converges as $`d0(\mathrm{})`$; $`i`$ labels the irreducible representation, $`\sigma `$ denotes spin. We have chosen the confining potential to be such, that at $`d0(\mathrm{})`$ the solutions of the double dot $`H_0`$ converge to the (shifted) Fock-Darwin functions, if properly symmetrized according to the representations of $`C_{2v}`$. These symmetrized functions will be denoted as $`g_i^{n,l,\sigma }`$, where (up to normalization) $$g_i^{n,l,\sigma }=(\mathrm{\Psi }_{n,l,\sigma }^d+D_i\mathrm{\Psi }_{n,l,\sigma }^d)+L_i(\mathrm{\Psi }_{n,l,\sigma }^d+D_i\mathrm{\Psi }_{n,l,\sigma }^d).$$ (19) The numbers $`D_i(L_i)`$ for different irreducible representations are in the Tab. 2. Generally, up to normalization, the exact solution can be written as a linear combination of any complete set of functions (we omit the spin index which is the same for all terms in the equation) $$\mathrm{\Gamma }_i^{ab}=\underset{n,l}{}\stackrel{~}{c}(n,l)g_i^{n,l}=g_i^{n_0,l_0}+\underset{n,l}{}c(n,l)g_i^{n,l}.$$ (20) The last equation indicates the fact, that for a function $`\mathrm{\Gamma }_i^{ab}`$ in the limit $`d0(\mathrm{})`$, there will be a dominant $`g`$-function in the sum with the numbers $`n_0,l_0`$ given by the level $`a(b)`$ and the coefficients $`c`$ for the other functions will converge to zero. We term the approximation $`c(n,l)=0`$ as a linear combination of single dot orbitals (LCSDO). Knowing the representations of the double dot Hamiltonian and the fact that Fock-Darwin functions form $`SO(2)`$ representations (reflecting the symmetry of single dot $`H_0`$) we can decompose all single dot levels into the double dot representations and thus formally construct the energy spectrum of a double dot using the symmetry considerations only. Following the standard technique for constructing such correlation diagrams (connecting states of the same representation and avoiding crossing of lines of the same representation) we arrive at the spectrum shown in Fig. 3. The ground state transforms by the symmetry operations according to $`\mathrm{\Gamma }_1`$ (identity), while the first excited state according to $`\mathrm{\Gamma }_2`$ ($`x`$). This is expected for the symmetric and antisymmetric states formed by single dot ground states. The symmetry structure of the higher excited states is important to understand spin-orbit coupling effects. Indeed, the spin-orbit terms couple two opposite spins according to certain selection rules. Since $`H_D`$, for example, transforms similarly to $`xy`$, it couples the ground state $`\mathrm{\Gamma }_1`$ with $`\mathrm{\Gamma }_2`$ and $`\mathrm{\Gamma }_4`$. In general, odd numbered representations can couple to even numbered representations. The same holds for $`H_{BR}`$ and $`H_{D3}`$. If we include either $`H_{BR}`$ or $`H_D`$ into the Hamiltonian, and consider spinors as the basis for a representation, the states would transform according to $`\mathrm{\Gamma }_5`$, the only irreducible representation of the double group of $`C_{2v}`$. The calculated numerical spectrum for our model structure is shown in Fig. 4. There is a nice qualitative correspondence with Fig. 3. In Fig. 4 by vertical bars we denote coupling through $`H_D`$ or $`H_{BR}`$ ($`|i|H_D|j|=|i|H_{BR}|j|`$). The couplings follow the selection rule described above. Since there are several level crossings in the lowest part of the spectrum, a question arises if spin hot spots are formed in the presence of spin-orbit coupling. It turns out, that there is no first-order level repulsion at the crossings due to the linear spin-orbit terms because the levels are not coupled by the linear terms, even though such couplings are allowed by symmetry. There are no spin hot spots due to the linear spin-orbit terms at zero magnetic field. For example $`\mathrm{\Gamma }_4^{11}`$ and $`\mathrm{\Gamma }_1^{21}`$ are not coupled by spin-orbit terms, and therefore their degeneracy (at $`2dl_050`$ nm) is not lifted by linear spin-orbit terms as we would expect from symmetry (actually, there is an anti-crossing which is of the order $`\alpha _{\mathrm{lin}}^3`$, instead of the expected $`\alpha _{\mathrm{lin}}`$). The cubic Dresselhaus term gives here (and also in other crossings that conform with the selection rules) a linear anti-crossing, as one expects. The absence of anti-crossings from the linear spin-orbit terms will be explained in the next section. Since our main goal here is to study the effects of spin-orbit coupling on the tunneling between the two dots, we first look at the tunneling for $`H_{SO}=0`$. The tunneling energy, $`\delta E_t/2`$, is given by the difference between the energies of the symmetric ground state and the asymmetric first excited orbital state: $`\delta E_t=E_AE_S`$. We compare the LCSDO approximation with our numerically exact calculations. The functions Eq. (19) are not orthogonal. If we introduce the overlap integrals between these functions (all the indices of a $`g`$-function are denoted by one letter here) by $`S_{ij}=i|j`$, and the Hamiltonian matrix elements $`H_{ij}=i|H|j`$, variational theory gives for the expansion coefficients $`c_i`$ of a double dot state $`\mathrm{\Gamma }=_ic_i|i`$ the generalized eigenvalue equation, $$\underset{j}{}H_{ij}c_j=E\underset{j}{}S_{ij}c_j.$$ (21) We will compute the energy of the two lowest orbital double dot states, $`\mathrm{\Gamma }_{1\sigma }^{00}\mathrm{\Gamma }_S^\sigma `$, $`\mathrm{\Gamma }_{2\sigma }^{10}\mathrm{\Gamma }_A^\sigma `$, which are a symmetric and an antisymmetric state with respect to $`x`$. The energies are denoted as $`E_S^{(0)},E_A^{(0)}`$, where index zero indicates the absence of spin-orbit coupling. We first use the LCSDO approximation suitable for the limit $`d\mathrm{}`$, that is we approximate $`\mathrm{\Gamma }_S^\sigma g_1^{0,0,\sigma }`$, and $`\mathrm{\Gamma }_A^\sigma g_2^{0,0,\sigma }`$. This means the basis in Eq. (21) consists of one function and the solution for the energy is $`E_i=H_{ii}/S_{ii}`$. For the considered states we obtain: $`E_S^{(0)}`$ $`=`$ $`2E_0{\displaystyle \frac{1+[12d/\sqrt{\pi }]e^{d^2}+d^2\mathrm{Erfc}(d)}{1+e^{d^2}}},`$ $`E_A^{(0)}`$ $`=`$ $`2E_0{\displaystyle \frac{1e^{d^2}+d^2\mathrm{Erfc}(d)}{1e^{d^2}}}.`$ (22) In the limit of large interdot separation (that is we make expansion in powers of $`e^{d^2}`$ and take the first term in this expansion as the leading order), $`E_{S(A)}^{(0)}E_0[2\pm (2/\sqrt{\pi })d\mathrm{exp}(d^2)]`$, where the minus (plus) sign is for $`E_S^{(0)}`$ ($`E_A^{(0)}`$). The tunneling energy then is $$\delta E_t^{(0)}E_0\frac{4}{\sqrt{\pi }}de^{d^2}.$$ (23) To understand this result, we get it once again by taking a two dimensional basis in Eq. (21) consisting of functions $`\mathrm{\Psi }_{0,0,\sigma }^d`$ and $`\mathrm{\Psi }_{0,0,\sigma }^d`$. In the limit $`d\mathrm{}`$ both diagonal elements of the matrix $`H`$ converge to the energy of the Fock-Darwin ground state ($`2E_0`$). Then the difference of the eigenvalues of $`H`$ is given by the off-diagonal matrix element, which is $`H_{12}=\mathrm{\Psi }_{0,0,\sigma }^d|H\mathrm{\Psi }_{0,0,\sigma }^d(\mathrm{\Psi }_{0,0,\sigma }^d\mathrm{\Psi }_{0,0,\sigma }^d)(\stackrel{}{r}=0)e^{d^2}`$. Any further refinement beyond LCSDO would not contribute significantly to the calculated $`\delta E_t^{(0)}`$. For example, for the ground state to go beyond LCSDO, we take the next excited function of the same symmetry and get the basis for Eq. (21) to be $`\{g_1^{0,0,\sigma },g_1^{0,1,\sigma }\}`$. The off-diagonal element in the matrix $`H`$ is $`H_{12}=g_1^{00}|H_0|g_1^{01}E_0P(d)e^{d^2}`$, where $`P(d)`$ is a polynomial in $`d`$. Since now $`H_{11}H_{22}`$ is of the order of $`E_0`$, the correction from non-diagonal terms will be of the order of $`(e^{d^2})^2`$ and will not change the leading order result. The numerical result, together with the analytical forms of $`\delta E_t^{(0)}`$ \[one using the complete expressions Eq. (22), the other for the limiting case of large $`d`$, Eq. (23)\], is plotted in Fig 5. The complete expression is in excellent agreement with the numerics, over the whole range of $`d`$, including the limit $`d0`$. As for the leading order expression, it becomes an excellent description for the tunneling energy at distances roughly twice the dot confinement length (40 nm in our case), as seen from the inset to Fig. 5. Using a LCSDO approximation $`g_1^{0,0,\sigma }`$ for the ground state is correct for both limits $`d0`$ and $`d\mathrm{}`$, because $`\mathrm{\Gamma }_{1\sigma }^{00}`$ converges to the single dot level 0 in both limits. Therefore it is reasonable to expect that the approximation will be equally good for the whole range of $`d`$. However, for the first excited state $`\mathrm{\Gamma }_2^{10}`$ the two limits go into different single dot levels and the proper LCSDO approximations for this state are $`g_2^{0,0,\sigma }`$ and $`g_2^{0,1,\sigma }`$ in the limit $`d\mathrm{}`$ and $`d0`$ respectively. However $`g_2^{0,0,\sigma }g_2^{0,1,\sigma }+g_4^{0,1,\sigma }`$ as $`d0`$ and thus for both ground and first excited states using a LCSDO approximation suitable for $`d\mathrm{}`$ gives good results in the whole range of $`d`$. We will see, that this will not be true for a non-zero magnetic field and describing $`\mathrm{\Gamma }_{2\sigma }^{10}`$ by $`g_2^{0,0,\sigma }`$ will give a much higher error. A remedy is to go beyond LCSDO for $`\mathrm{\Gamma }_2^{10}`$, for example by taking the base for Eq. (21) to include two $`g`$-functions, each correct in one of the limits $`d0`$, $`d\mathrm{}`$. We will not present the results of such computations since formulas become more complicated without giving better understanding. Higher orbital states can be treated similarly. Starting from level 2 there are more functions of the same representation in one single dot level, therefore the basis for the Hamiltonian Eq. (21) giving the leading order must contain more that one function. ### IV.2 Corrections to energy from spin-orbit coupling in zero magnetic field. When we add $`H_{SO}`$ to $`H_0`$, the structure of the corrections to the energies of the two lowest states up to the second order in spin-orbit couplings can be expressed as $$E_i^{(2)}=𝒜_i(\alpha _D^2+\alpha _{BR}^2)_i\alpha _{D3}^2+𝒞_i\alpha _D\alpha _{D3},$$ (24) where $`i`$ is either $`S`$ or $`A`$. The coefficients $`𝒜,,`$ and $`𝒞`$ are positive for all values of the interdot distance. The differences $`𝒜_A𝒜_S,\mathrm{}`$ approach zero as $`d\mathrm{}`$. We will argue below that $`𝒜_S=𝒜_A=1/2`$ with the exception of a very small interdot distance (less than 1 nm). There are thus no contributions from the linear spin-orbit couplings to $`\delta E_t`$ in the second order. Only the cubic Dresselhaus term contributes, either by itself or in combination with the linear Dresselhaus term. Spin-dependent tunneling is greatly inhibited. Numerical calculation of the corrections to $`\delta E_t`$ from spin-orbit couplings are shown in Fig 6. The dominant correction is the mixed $`D`$-$`D3`$ term, followed by $`D3`$-$`D3`$. These are the only second order corrections. For GaAs, and our model geometry, these corrections are about 4 and 5 orders of magnitude lower than $`\delta E_t^{(0)}`$. The corrections, when only linear spin-orbit terms are present, are much smaller since they are of the fourth order. The dramatic enhancement of the corrections from linear spin-orbit terms close to $`d=0`$ is due to the transition from coupled to single dots. We will explore this region in more detail later. We will first show how naive approaches to calculating spin-orbit contributions to tunneling fail to explain the above results. We use the example of the linear Dresselhaus term. The simplest way to include this term is to begin with the two lowest orbital states (that are four states including spin), $`g_1^{0,0,\sigma }`$ and $`g_2^{0,0,\sigma }`$. Because of the time reversal symmetry the resulting $`4\times 4`$ matrices $`H`$ and $`S`$ from Eq.(21) block diagonalize into two equal $`2\times 2`$ matrices with elements $`H_{11}=E_S^{(0)}`$, $`H_{22}=E_A^{(0)}`$, and $`H_{12}=g_1^{0,0,}|H_D|g_2^{0,0,}=iE_0\alpha _Dde^{d^2}`$; $`S`$ is the unit matrix now, since the two states are orthogonal due to symmetry. Using the large $`d`$ limit for $`\delta E_t^{(0)}`$, Eq. (23), we obtain the perturbed energies $`E_{S(A)}=2E_0\pm E_0\sqrt{4/\pi +\alpha _D^2}de^{d^2}`$ with the minus (plus) sign for $`S`$ ($`A`$). In the second order of $`\alpha _D`$ the symmetric and antisymmetric level energies have opposite contributions, giving $`\delta E_t[(4E_0/\sqrt{\pi })+(E_0\sqrt{\pi }/2)\alpha _D^2]d\mathrm{exp}(d^2)`$, in contrast to the numerical results where there is no dependence on $`\alpha _D^2`$ in the second order. Enlarging the basis by the first excited orbital states (all together 12 states including spin), that is, include $`g_i^{0,1,\sigma }`$, the symmetric and antisymmetric level energies will have the correct limit for the spin-orbit contributions, $`E_0\alpha _D^2/2`$, at $`d\mathrm{}`$. At finite values of $`d`$ the difference from this limit value is less than 2%. That means there are still terms of order $`\alpha _D^2`$ in $`\delta E_t`$. Could using a renormalized basis help? We could, for example, use symmetrized states of the separated dots that include spin-orbit terms. It is not difficult to see that this would not work either: the perturbed single dot ground state, for example, contains the spin admixture from the first excited orbital states. This is then similar to using the 12 state basis in the variational approach. From the previous example one can see that to get a correct spin-orbit contribution to the energy of a state, it is not enough to include just a few terms in the sum in Eq. (9). Instead we employ the operators $`H^{op}`$ given in Eqs. (11,12). To get a contribution for a particular state, say $`|i`$, we apply the Löwdin perturbation theory.Löwdin (1951) For this one has to identify states $`|j`$ which are degenerate with $`|i`$ with respect to the perturbation $`H_{SO}`$ and which have to be treated exactly. The rest of the states can be treated perturbatively. The condition for a degeneracy of two states can be taken as $`|E_i^{(0)}E_j^{(0)}|\alpha _{SO},\alpha _{\mathrm{lin}}`$, when one considers linear and cubic terms respectively. The finite set of the degenerate states will be denoted by $`𝒩`$. The effective Hamiltonian $`H^{\mathrm{eff}}`$ acting in $`𝒩`$ is $`H_{ij}^{\mathrm{eff}}=(H_0+H_{SO})_{ij}+`$ $`+{\displaystyle \frac{1}{2}}{\displaystyle \underset{k𝒩}{}}[{\displaystyle \frac{(H_{SO})_{ik}(H_{SO})_{kj}}{E_i^{(0)}E_k^{(0)}}}+{\displaystyle \frac{(H_{SO})_{ik}(H_{SO})_{kj}}{E_j^{(0)}E_k^{(0)}}}].`$ (25) For the example of the linear Dresselhaus term, we can now use Eq. (10) and (13) to obtain $$H_{ij}^{\mathrm{eff}}=(H_0+H_D)_{ij}\frac{1}{2}\alpha _D^2E_0\left(1\sigma _zl_z\right)_{ij}+R_{ij},$$ (26) where $$R_{ij}=\frac{1}{2}i|H_DP_𝒩H_D^{op}H_D^{op}P_𝒩H_D|j.$$ (27) First we note that existence of the operator $`H_D^{op}`$ means that the coupling through $`H_D`$ between any two states is always much smaller then the difference of the unperturbed energies of these two states, since $`(H_D)_{ij}=(E_i^{(0)}E_j^{(0)})(H_D^{op})_{ij}(E_i^{(0)}E_j^{(0)})\alpha _D`$. Then one can partially diagonalize the effective Hamiltonian to eliminate the off-diagonal $`H_D`$ terms. It turns out, that this leads to a cancellation of the terms $`H_D`$ and $`R`$. The effective Hamiltonian is then $$H_{ij}^{\mathrm{eff}}=(H_0)_{ij}\frac{1}{2}\alpha _D^2E_0\left(1\sigma _zl_z\right)_{ij}.$$ (28) This completes the way to get Eq. (16) using Löwdin perturbation theory. There are no linear effects on the double dot energy spectrum from linear spin-orbit terms, which explains the absence of spin hot spots even though symmetry allows that. The spin-orbit interaction can influence the energy only through the operator $`l_z`$, which is of the representation $`\mathrm{\Gamma }_3`$, from where we get selection rule–the allowed coupling is between functions of representations $`\mathrm{\Gamma }_1`$$`\mathrm{\Gamma }_3`$ and $`\mathrm{\Gamma }_2`$$`\mathrm{\Gamma }_4`$. Looking at Fig. 4, accidental degeneracies of states with such representations are not present in the lower part of the spectrum. The crossing of $`\mathrm{\Gamma }_1^{21}`$ with $`\mathrm{\Gamma }_4^{11}`$ considered in the discussion to Fig. 4 also does not follow the selection rule, hence why the anti-crossing is of the third order. From the selection rules one can immediately see that also the expectation value of $`l_z`$ is zero in any state. This result is more general and holds also if the symmetry of the potential is lower (or none), since it follows from the fact that the Hamiltonian $`H_0`$ is real, so one can choose eigenfunctions to be real. Then the expectation value of any imaginary operator, such as $`l_z`$, must vanish. We conclude, that apart from degeneracies following from the single dot \[that is limits $`d0(\mathrm{})`$\] and possible accidental degeneracies respecting the selection rule, double dot states are described by an effective Hamiltonian $$H_{ii}^{\mathrm{eff}}=E_i^{(0)}\frac{1}{2}E_0\alpha _D^2.$$ (29) Particularly, the energies of the two lowest states are given by this equation, with an exception for the state $`\mathrm{\Gamma }_A`$ in the region of small $`d`$ where it is coupled to $`\mathrm{\Gamma }_4^{11}`$ through $`l_z`$ and we have to describe it here by a $`2\times 2`$ effective Hamiltonian. An illustration of the $`l_z`$ influence on the spectrum is in Fig. 7, where the linear Dresselhaus spin-orbit contribution to the energy for several states as a function of the interdot distance is shown. One can see at what interdot distances the $`l_z`$ operator causes the qualitative change between the double dot case (where the functions are characterized by a definite representation $`\mathrm{\Gamma }_i`$ and the energy contribution from the spin-orbit is a uniform shift) and the single dot case (where the functions are numbered according to the orbital momentum and the spin-orbit contribution to the energy depends on $`\sigma _zl_z`$). This happens when $`E_0\alpha _D^2`$ is comparable to the energy difference of the nearly degenerate states. If the criterion for the coupling between the dots is the constant contribution, $`\alpha _D^2E_0/2`$, to the energy, then the double dot region, as far as the spin-orbit coupling is concerned, is between 1 to 100 nm, that is up to 5 times of the confinement length of 20 nm. As an example, for the function $`\mathrm{\Gamma }_4^{11}`$ the coupling in the effective Hamiltonian through $`l_z`$ to $`\mathrm{\Gamma }_2^{31}`$ is equal to the unperturbed energy difference if $`\alpha _D^2d^3e^{d^2}`$, giving $`d3`$, corresponding to the interdot distance of $`6l_0`$. Due to the exponential, this result is insensitive to $`\alpha _D`$. The Bychkov-Rashba term can be treated analogously. The effective Hamiltonian is $`H_{ij}^{\mathrm{eff}}=[H_0(\alpha _{BR}^2/2)E_0(1+\sigma _zl_z)]_{ij}`$. The absence of a linear influence on the energy was based on the existence of $`H_D^{op}`$. Since we found a case where $`H_{D3}`$ causes linear anti-crossing (see discussion to Fig. 4), it follows that $`H_{D3}^{op}`$ can not exist for our double dot potential. However, if one approximates $`E_iE_kE_jE_k`$ in (IV.2), one can use $`H_D^{op}`$ to simplify the mixed $`D`$-$`D3`$ correction. This, according to Fig. 6, is the dominant spin-orbit correction for the tunneling energy $`\delta E_t`$. One gets an analogous expression as Eq. (26), where the needed commutator is stated in Eq. (15). Concluding, if we neglect the mixed $`D3`$-$`D3`$ term, we can write the spin-orbit contribution to the energy for the lowest orbital states to be ($`i=S,A`$) $$\delta E_i^{SO}=E_0(\alpha _D^2+\alpha _{BR}^2)/2+\alpha _D\alpha _{D3}E_0l_0^2(\text{k}^2)_{ii}.$$ (30) One note to the eigenfunctions: The matrix elements of the effective Hamiltonian are computed using the eigenfunctions of $`H_0`$. But the functions that correspond to the solutions are transformed by the same unitary transformation that leads from $`H_0`$ to $`H^{\mathrm{eff}}`$. The sum rule can be used also here to express the influence of $`H_{\mathrm{lin}}`$ on the eigenfunctions of $`H_0`$. If $`H_0\mathrm{\Gamma }_i=E_i\mathrm{\Gamma }_i`$, the eigenfunctions corresponding to the effective Hamiltonian, Eq. (26), are $$\delta \mathrm{\Gamma }_i=\underset{j𝒩}{}\frac{(H_{\mathrm{lin}})_{ji}}{E_iE_j}\mathrm{\Gamma }_j=(𝕀P_𝒩)H_{\mathrm{lin}}^{op}\mathrm{\Gamma }_i.$$ (31) Partial diagonalization of the effective Hamiltonian, to go from Eq. (26) to Eq. (28), means we finish the unitary transformation completely and get $`\stackrel{~}{\mathrm{\Gamma }}_i^\sigma =\mathrm{\Gamma }_i^\sigma H_{\mathrm{lin}}^{op}\mathrm{\Gamma }_i^\sigma `$ for the eigenfunctions corresponding to the effective Hamiltonian, Eq. (28). ### IV.3 Finite magnetic field. The presence of a magnetic field lowers the symmetry of the Hamiltonian without spin-orbit terms. The only nontrivial symmetry operator is the inversion $`I`$ (see Tab. 2). As a consequence the double dot states fall into two groups (representations of $`C_2`$): $`\mathrm{\Gamma }_1`$ and $`\mathrm{\Gamma }_3`$ become $`\mathrm{\Gamma }_S`$ (symmetric under $`I`$) and $`\mathrm{\Gamma }_2`$ and $`\mathrm{\Gamma }_4`$ become $`\mathrm{\Gamma }_A`$ (antisymmetric under $`I`$). Symmetrized functions $`g_i^{n,l,\sigma }`$ now are $`g_i^{n,l,\sigma }=\mathrm{\Psi }_{n,l,\sigma }^d+D_i\mathrm{\Psi }_{n,l,\sigma }^d,`$ (32) where the irreducible states $`i=S`$ and $`A`$, while the permutation coefficients $`D_S=D_A=1`$. The shifted single-dot wave functions acquire a phase: $`\mathrm{\Psi }_{n,l,\sigma }^{\pm d}(x,y)=\mathrm{\Psi }_{n,l,\sigma }(x\pm l_0d,y)e^{\pm idl_0\theta y/l_B^2},`$ (33) depending on which dot they are located. The double dot energy spectrum of $`H_0`$ as a function of magnetic field is shown in Fig. 8 for the interdot distance of 50 nm. Indicated are two crossings and one anti-crossing induced by magnetic field. The first crossing is between $`\mathrm{\Gamma }_1`$ and $`\mathrm{\Gamma }_2`$ (notation from the $`B=0`$ case). These two states have opposite spins so they are not coupled and there is no level repulsion here. (We will see in the next section that spin-orbit coupling will induce anti-crossing in this case.) The second crossing is between $`\mathrm{\Gamma }_2`$ and $`\mathrm{\Gamma }_3`$, which behave differently under $`I`$ and so they are not coupled by magnetic field. The actual anti-crossing is between $`\mathrm{\Gamma }_2`$ and $`\mathrm{\Gamma }_4`$, which are both antisymmetric under $`I`$. This is an example of anti-crossing induced by magnetic field. In analogy with Eq. (22) we derive analytical expressions for the energies of the lowest symmetric and antisymmetric states in the presence of magnetic field using the LCSDO approximation: $`E_S^{(0)}`$ $`=`$ $`{\displaystyle \frac{2E_0}{\eta ^2}}({\displaystyle \frac{1+[1d\eta (1\theta ^2)/\sqrt{\pi }]e^{(d\eta )^2(1+\theta ^2)}}{1+e^{(d\eta )^2(1+\theta ^2)}}}`$ $`{\displaystyle \frac{d\eta (1\theta ^2)[e^{(d\eta )^2}/\sqrt{\pi }d\eta \mathrm{Erfc}(d\eta )]}{1+e^{(d\eta )^2(1+\theta ^2)}}}),`$ $`E_A^{(0)}`$ $`=`$ $`{\displaystyle \frac{2E_0}{\eta ^2}}({\displaystyle \frac{1[1d\eta (1\theta ^2)/\sqrt{\pi }]e^{(d\eta )^2(1+\theta ^2)}}{1e^{(d\eta )^2(1+\theta ^2)}}}`$ (34) $`{\displaystyle \frac{d\eta (1\theta ^2)[e^{(d\eta )^2}/\sqrt{\pi }d\eta \mathrm{Erfc}(d\eta )]}{1e^{(d\eta )^2(1+\theta ^2)}}}).`$ Here $`\eta =l_0/l_B=(1\theta ^2)^{1/4}`$. In the limit $`d\mathrm{}`$, we can then deduce the tunneling energy in the leading order to be $`\delta E_t^{(0)}=E_0{\displaystyle \frac{4}{\sqrt{\pi }}}(1\theta ^2)^{5/4}e^{d^2(1+\theta ^2)/\sqrt{1\theta ^2}}.`$ (35) If $`\theta =0`$, the above expressions reduce to Eq. (22,23). On the other hand, if $`B\mathrm{}`$, then $`\delta E_t^{(0)}B^{5/2}e^{B/B_0}`$. Magnetic field suppresses $`\delta E_t^{(0)}`$ by suppressing overlap integrals $`\mathrm{\Psi }^d|H|\mathrm{\Psi }^d`$. There are three different effects that magnetic field introduces. First, the wave functions are squeezed by the confinement provided by the vector potential. The natural confinement length is $`l_B=l_0(1\theta ^2)^{1/4}`$, present in the exponential decay factors. Second, the gauge phase produces factors $`(1+\theta ^2)`$ in the exponents of the scalar products $`\mathrm{\Psi }^d|\mathrm{\Psi }^d`$. Third, as $`B`$ increases, the confinement potential $`V_C`$ becomes less important compared to the confinement of the magnetic vector potential. This gives rise to the factor $`(1\theta ^2)`$. Note that in the limit $`B\mathrm{}`$, $`\mathrm{\Psi }_{n,l,\sigma }^d`$ is an eigenstate of $`H_0`$ for any $`d`$. For reasons explained in last paragraph in the Sec. 4, Eqs. (34) are correct in the limit $`d\mathrm{}`$. As $`d0`$, only $`E_S^{(0)}`$ is correct. The limit value of $`E_A^{(0)}(d=0)/E_0`$ is $`4/(1+\theta ^2)`$, instead of the exact value of $`2(2\theta )`$. At a finite magnetic field we have also a new term in the Hamiltonian, the Zeeman term. Since it commutes with $`H_0`$ the only consequence of this term is a shift of the energy of the states by a value $`\sigma \mu _BB`$ according to their spin $`\sigma `$. Therefore it introduces new crossings of the states with opposite spin. An example of this can be seen in Fig. 9, where we plot energies of the four lowest states in the region where the Zeeman shift is comparable to the energy differences of the considered states. ### IV.4 Effective spin-orbit Hamiltonian We now study the influence of spin-orbit coupling on the spectrum of double dots in a finite magnetic field. We will see that spin-orbit terms lead to new spin hot spots even at magnetic fields of the order of 1 T, and that linear spin-orbit terms will influence tunneling in the second order. Although the presence of the Zeeman term complicates the analysis of the perturbation theory using operators $`H^{op}`$, one can still apply the previously developed formalism if the Zeeman term is treated as a part of perturbation. (For a harmonic potential describing single dots, operators $`H_{\mathrm{lin}}^{op}`$ have been derivedValin-Rodriguez (2004) for the case of finite magnetic field, so that the Zeeman term can be included into $`H_0`$). Up to the second order in the perturbation couplings (being now $`\alpha _{SO}`$ and $`\alpha _Z`$), there is no coupled Zeeman-spin-orbit term. This means that in the effective Hamiltonians $`H^{\mathrm{eff}}`$ that we already derived for the case of zero magnetic field, the Zeeman term appears as a shift of the energies on the diagonal without bringing any new couplings (non-diagonal terms). But an important consequence is that the shift can change the number of states we have to include into the basis where the effective Hamiltonian acts, because their energy difference to the considered state is comparable to the spin-orbit coupling. First, in analogy with Eq. (29), if the energy of a state is far enough from others, we can consider the basis to consist of one term only and the spin-orbit correction to the energy of state $`|i`$ is $`\delta E_i^{SO}`$ $`=`$ $`E_0{\displaystyle \frac{\alpha _D^2}{2}}(1\sigma \overline{L_z})E_0{\displaystyle \frac{\alpha _{BR}^2}{2}}(1+\sigma \overline{L_z})`$ (36) $`+\overline{[H_D^{op},H_{D3}]},`$ where the averaging means the expectation value in the state $`|i`$ and $`\sigma `$ is the spin of the state. Since the presence of magnetic field lowers the symmetry, the last commutator, \[Eq. (15)\], can not be simplified according to the symmetry as was the case before in Eq. (30), and, more important, we no longer have $`\overline{L_z}=0`$. As a result, there are now corrections to the tunneling that are of the second order in the linear spin-orbit couplings. These corrections depend on $`\alpha _{}^{(2)}\alpha _D^2\alpha _{BR}^2`$. Second, we look how the energies of the four lowest states are changed, using again the example of the linear Dresselhaus term. They are plotted in Fig. 9. Here in the main figure one can see the shift caused by the Zeeman term and the anti-crossing induced by the spin-orbit coupling is magnified in the inset. The anti-crossing states are $`\mathrm{\Gamma }_S^{}`$ and $`\mathrm{\Gamma }_A^{}`$. In the case of zero magnetic field we described each of the four basis states by Eq. (29). Now, in principle, we have to describe them by a $`4\times 4`$ effective Hamiltonian Eq. (26). Due to symmetry we can simplify this Hamiltonian into two $`2\times 2`$ Hamiltonians, $`H_1^{\mathrm{eff}},H_2^{\mathrm{eff}}`$, acting in the bases $`\mathrm{\Gamma }_S^{},\mathrm{\Gamma }_A^{}`$ and $`\mathrm{\Gamma }_S^{},\mathrm{\Gamma }_A^{}`$ respectively. The four components of the effective Hamiltonian matrix are $`(H^{\mathrm{eff}})_{11}`$ $`=`$ $`E_S^{(0)}E_0{\displaystyle \frac{\alpha _D^2}{2}}(1\sigma (L_z)_{11})\sigma \mu _BBR_{11}`$ $`(H^{\mathrm{eff}})_{22}`$ $`=`$ $`E_A^{(0)}E_0{\displaystyle \frac{\alpha _D^2}{2}}(1+\sigma (L_z)_{22})+\sigma \mu _BB+R_{11}`$ $`(H^{\mathrm{eff}})_{12}`$ $`=`$ $`(H^{\mathrm{eff}})_{21}^{}=(H_D)_{12}`$ (37) where $`\sigma =1`$ for $`H_1^{\mathrm{eff}}`$ and $`\sigma =1`$ for $`H_2^{\mathrm{eff}}`$, while indices 1,2 denote the first and the second term in the corresponding basis. Comparing to the case of zero magnetic field the Zeeman term increases the difference of the diagonal elements in $`H_1^{\mathrm{eff}}`$ and decreases them in $`H_2^{\mathrm{eff}}`$. The ground and the fourth excited states which are described by $`H_1^{\mathrm{eff}}`$ stay isolated, and we can do the perturbative diagonalization to get rid of the off-diagonals. The energy of the two states is then up to the second order in the spin-orbit coupling accurately described by Eq. (36). Concerning the states $`\mathrm{\Gamma }_S^{}`$ and $`\mathrm{\Gamma }_A^{}`$, there is a region in the interdot distance of a few nanometers, where the two states must be described by the two dimensional $`H_2^{\mathrm{eff}}`$ to account for the anti-crossing, which is caused by the the matrix element $`\mathrm{\Gamma }_S^{}|H_D|\mathrm{\Gamma }_A^{}`$. LCSDO gives for this element a result correct only in the order of magnitude. This is because even in the limit $`d\mathrm{}`$ this matrix element is of the same order as the neglected coefficients $`c(n,l)`$ in the LCSDO approximation, Eq. (20). The spin-orbit corrections to the energies from $`H_D`$ for the four lowest states as functions of the interdot distance are in Fig. 10. Also shown are analytical values computed by Eq. (36), that is, ignoring anti-crossing. The scale implies that the corrections are of the second order in $`\alpha _D`$ and for the states $`\mathrm{\Gamma }_S^{}`$ and $`\mathrm{\Gamma }_A^{}`$ are enhanced in the anti-crossing region. The incorrect values at the limit $`d0`$ for functions $`\mathrm{\Gamma }_A^\sigma `$ are because of reasons explained in the last paragraphs in Sec. 4. ### IV.5 Spin-orbit corrections to the effective $`g`$-factor and tunneling frequency We next analyze spin-orbit corrections to the $`g`$-factor, $`\delta g[\delta E(\mathrm{\Gamma }_S^{})\delta E(\mathrm{\Gamma }_S^{})]/\mu _BB`$, that characterizes the energy cost of a spin flip in the ground state, or the frequency of a spin precession. Another kind of oscillation is electron tunneling, when electron oscillates between the left and the right dot without changing its spin. The frequency of this oscillation, $`\delta E_t/2\mathrm{}`$, is given by the energy difference $`\delta E_t=E(\mathrm{\Gamma }_A^{})E(\mathrm{\Gamma }_S^{})`$. Corrections to this energy difference induced by the spin-orbit interaction are denoted in this paper as $`\delta E_t^{SO}`$. First, we take a look at the spin-orbit corrections to the $`g`$-factor. Contributions in the second order of the spin-orbit couplings are shown in Fig. 11, as functions of magnetic field at a constant interdot distance. The spin-orbit contribution to the $`g`$-factor in the double dot case has the same qualitative dependence on the magnetic field as in the single dot case (see discussion to Fig. 2). However, at finite interdot distances, there is an enhancing effect on the spin-orbit contributions. This can be seen in Fig. 10, where at a certain magnetic field, the spin-orbit contribution to the $`g`$-factor is enhanced for a finite $`d`$ compared to the case of $`d=0(\mathrm{})`$. We found numerically, that the enhancement can be up to 50% of the value of the correction in $`d=0`$ at magnetic fields of the order of 1 T. At the anti-crossing the spin-orbit contributions show cusps. At magnetic fields bellow the anti-crossing, the dominant spin-orbit contribution is $`D`$-$`D`$ which reduces the conduction band $`g`$-factor by several percent. Contributions $`D`$-$`D3`$ and $`BR`$-$`BR`$ are one order of magnitude smaller. Using Eq. (36), that is ignoring the anti-crossing, we get for the contribution from the linear spin-orbit terms $$\delta g(\mathrm{lin}\mathrm{lin})=\frac{E_0}{\mu _BB}\alpha _{}^{(2)}\overline{L_z},$$ (38) where, in the limit $`d\mathrm{}`$, $`\overline{L_z}\theta [1+(d/\eta )^2e^{(d\eta )^2(1+\theta ^2)}].`$ (39) From Fig. 11 one can see that the analytical result agrees with numerics. Finally, we look at the tunneling energy in the presence of both magnetic field and spin-orbit couplings. The spin-orbit corrections, as a function of magnetic field, are shown in Fig. 12. At zero magnetic field there is no contribution from the linear terms (result of the section IV.2) and the dominant contribution is $`D`$-$`D3`$. Similarly to $`\delta E_t^{(0)}`$, the corrections decay exponentially with increasing magnetic field. Anti-crossing strongly influences the tunneling energy. Using LCSDO for $`d\mathrm{}`$ we obtain in the leading order $$\delta E_t^{\mathrm{lin}}=E_0\alpha ^{(2)}\theta (d/\eta )^2e^{(d\eta )^2(1+\theta ^2)}.$$ (40) This analytical formula underestimates the corrections from the linear spin-orbit terms by a factor of $`3`$. Nevertheless the analytical expression for $`D`$$`D3`$ is reasonably good. In the magnetic field below anti-crossing, the relative change of the tunneling energy stemming from the spin-orbit terms is of order $`10^3`$. ### IV.6 Tunneling Hamiltonian We now use our results to describe the influence of the spin-orbit interaction on the lower part of the spectrum. We restrict our Hilbert space on the four lowest states $`\mathrm{\Gamma }_{S(A)}^\sigma `$, the eigenstates of the total double dot Hamiltonian. Because of the transformation Eq. (31) these four states have neither definite symmetry with respect to inversion $`I`$, nor a definite spin in z-direction. In this section we will denote them as spin ‘up’ and spin ‘down’ states. For description of a transport through the double dot it is useful to define the following left and right localized functions $$L^\sigma (R^\sigma )=\frac{1}{\sqrt{2}}(\mathrm{\Gamma }_S^\sigma \pm \mathrm{\Gamma }_A^\sigma ),$$ (41) where plus and minus holds for $`L`$ and $`R`$, respectively. In the limit $`d\mathrm{}`$ these functions converge to single dot solutions localized in the left and right dot. The effective Hamiltonian of our system in the second quantization formalism is $$H=\underset{\sigma =,}{}E^\sigma (n_{L\sigma }+n_{R\sigma })(\delta E_t^\sigma /2)(a_{L\sigma }^{}a_{R\sigma }+a_{R\sigma }^{}a_{L\sigma }),$$ (42) where $`E^\sigma =(E_S^\sigma +E_A^\sigma )/2`$, $`\delta E_t^\sigma =E_A^\sigma E_S^\sigma `$, while $`a^{}`$, and $`a`$ are creation and annihilation operators, and $`n=a^{}a`$. We can get both localized and spin-pure states if we diagonalize $`\sigma _z`$ in a chosen basis. For example, taking $`L^{}`$ and $`L^{}`$, we get $`L^{\text{pure}}(L^{}+oL^{})`$ and $`L^{\text{pure}}(L^{}o^+L^{})`$, up to normalization $`(1|o|^2/2)`$. That is, the left pure spin state is a linear superposition of both left states with spin ‘up’ and ‘down’. The coefficient $`o`$ is proportional to $`\alpha _{SO}`$. In the following we are interested in the time evolution of localized states given by Hamiltonian Eq. (42). First we note, that due to the non-diagonal terms, the electron which is in a localized state will tunnel into the other localized state after the tunneling time $`t_{\mathrm{tun}}^\sigma =h/\delta E_t^\sigma `$, resulting in coherent oscillations. For our parameters $`t_{\mathrm{tun}}1`$ ps. In the Hamiltonian there is no mixing between spin ‘up’ and ‘down’ states. However, there will be mixing (or spin-flip) if we work with localized pure spin states. Electron being originally in $`L^{\text{pure}}`$ will, after the tunneling time $`t_{\mathrm{tun}}^{}`$, contain $`R^{\text{pure}}`$ with the probability amplitude $$c=io\pi (\delta E_t^{}\delta E_t^{})/2\delta E_t^{},$$ (43) assuming that the difference in $`\delta E_t`$ for different spins is much smaller that $`\delta E_t`$ itself. In the case of zero magnetic field, because of Kramer’s degeneracy, the tunneling frequencies are the same for both spin orientations. Then whatever is the initial combination of spin ‘up’ and ‘down’ (let it be, for example, $`L^{\text{pure}}`$), during the time evolution (oscillations) there will be no relative change in the coefficients in this linear combination. Therefore spin-orbit coupling does not lead to spin-flipping and $`c=0`$ in Eq. (43). In a finite magnetic field, the tunneling frequency for spin ‘up’ and ‘down’ are different. The difference is caused only by spin-orbit terms, and is of order $`\alpha _{}^{(2)}=\alpha _D^2\alpha _{BR}^2`$. Equation (36) shows, that the contribution to $`\delta E_t^{}`$ from the linear spin-orbit terms is opposite that of $`\delta E_t^{}`$ and therefore their difference is twice the expression in Eq. (40). We conclude, that spin-flip during tunneling induced by spin-orbit coupling is proportional to the third power in spin-orbit couplings and depends linearly on the magnetic field if the magnetic field is small ($`c\alpha _{\mathrm{lin}}^3\theta `$). The different tunneling frequency can be exploited for separation of different spin states in a homogeneous magnetic field. Starting with some combination of ‘up’ and ‘down’ states localized in one dot, after time $`t_{\mathrm{sep}}=h/(\delta E_t^{}\delta E_t^{})=t_{\mathrm{tun}}\delta E_t^{}/(\delta E_t^{}\delta E_t^{})`$ the part with spin ‘up’ will be localized in the left, and the spin ‘down’ will be in the right dot. From Fig. 12 one can see that one needs about $`10^3`$ coherent oscillations to get the spatial spin separation. Therefore the decoherence time must be longer in order to observe this effect experimentally. We note that the separated states will not be pure spin states, but will contain a small (linearly proportional to $`\alpha _{SO}`$) admixture of opposite pure spin states. ## V Conclusions We have performed numerically exact calculation of the spectrum of a single electron localized by a confining potential in single and double GaAs quantum dots. We have studied the influence of the spin-orbit terms, namely the Bychkov-Rashba and the linear and cubic Dresselhaus terms, on the energy spectrum. In the single dot case we have elaborated on previous results and shown that the spin-orbit interaction has three principal effects on the spectrum: first, the interaction shifts the energy by a value proportional to the second order in the spin-orbit couplings, second, it lifts the degeneracy at zero magnetic field, and, third, the Bychkov-Rashba term gives rise to spin hot spots at finite magnetic fields. In the double dot case we have addressed the symmetries of the Hamiltonian. For zero magnetic field without spin-orbit terms we have constructed the correlation diagram, between single and double dot states, of the spectrum. We have used properly symmetrized linear combination of shifted single dot solutions as an approximation for a double dot solution and found that for the four lowest states it gives a good approximation for the energy. As for the contributions to the energy from the linear spin-orbit terms, we have found that in zero magnetic field a typical feature of a double dot is a uniform shift of the energy proportional to the second order in the coupling strengths without any dependence on the interdot distance. This is true also if the potential has lower (or none) symmetry (for example biased dots). Therefore, in zero magnetic field, there is no influence on the tunneling frequency up to the second order in the linear spin-orbit couplings and the dominant contribution comes from the mixed linear and cubic Dresselhaus second order term. We found, that spin hot spots in zero magnetic field exist in the double dot, but are solely due to the cubic Dresselhaus term. This means also, that for our potential, for the cubic Dresselhaus term there can not exist a unitary transformation to eliminate its contribution in the first order. The effective $`g`$-factor, on the other hand, is influenced by the second order linear spin-orbit couplings even at $`B0`$, so the dominant contribution here is the linear Dresselhaus term for GaAs. In finite magnetic fields the uniform shift does not hold any more and there is a contribution to the tunneling frequency in the second order of the linear spin-orbit couplings. We have derived an effective Hamiltonian, using Löwdin’s perturbation theory, with which analytical results up to the second order in perturbations (Zeeman and spin-orbit terms with the exception of cubic Dresselhaus-cubic Dresselhaus contribution) can be obtained provided one has exact solutions of the double dot Hamiltonian without Zeeman and spin-orbit terms. From this effective Hamiltonian we have derived the uniform shift in zero magnetic field. In a finite magnetic field we used linear combinations of single dot solutions to obtain analytical expressions for the spin-orbit contributions to the energy for the four lowest states. We have analyzed them as functions of the interdot distance and magnetic field and compared them with exact numerical values. The spin-orbit relative contribution to the $`g`$-factor and the tunneling frequency is of the order of $`10^2`$ and $`10^3`$, respectively. Due to the degeneracy of the energy spectrum at large interdot distance the spin hot spots exist also at smaller magnetic fields compared to the single dot case. As an application of our results we have constructed an effective Hamiltonian acting in a restricted Hilbert space of four states–electron with spin up and down (these are effective spins in the presence of spin-orbit coupling) localized on either dot. Effectively, there is only spin-conserving tunneling between the localized states, no spin-flip tunneling. In zero magnetic field the spin-orbit interaction does not significantly influence the tunneling frequency, nor it implies spin-flip tunneling. In finite magnetic fields the tunneling frequency is spin dependent, the difference being of second order in linear spin-orbit terms. This leads to a spin flip amplitude proportional to the third power in spin-orbit couplings (it is linear in magnetic field). We propose to use this difference of the tunnelings to spatially separate electron spin in homogeneous magnetic field. ###### Acknowledgements. We thank Ulrich Hohenester for usefull discussions. This work was supported by the US ONR.
warning/0506/astro-ph0506602.html
ar5iv
text
# Detection of gravity waves by phase modulation of the light from a distant star ## I Introduction The detection of gravity waves (GW) has stimulated a lot of interest for decades. There are two major GW detection concepts: acoustic and interferometric detection. The acoustic method deals with a resonance response of massive elastic bodies on GW excitations. Historically the acoustic method was proposed first by J. Weber weber where he suggested to use long and narrow elastic cylinders as GW antennas. Although a significant progress has been achieved in fabrication and increasing sensitivity of such type of detectors amaldi\_1 ; amaldi\_2 ; astone the interpretation of obtained data is still far to claim undoubtedly the detection of GW. On the other hand a considerable attention has been shifted recently to more promising interferometric detection methods. The interferometric gravitational-wave detector like Laser Interferometric Gravitational Wave Observatory (LIGO) and VIRGO vogt ; bradaschia represents a Michelson interferometer with a laser beam split between two perpendicular arms of interferometer. The principles of operation of such type of detectors are reviewed in Refs forward ; thorne ; hough ; drever ; meers . The action of gravitational waves on an interferometer can be presented as relative deformation of both interferometer arms. A gravitational wave with dimensionless amplitude $`h`$ induces the opposite length changes $`\delta l/l=1/2h\mathrm{cos}\mathrm{\Omega }t`$ in each arm of the Michelson interferometer, where $`l`$ is the length of the arm, $`\mathrm{\Omega }`$ is the gravitational wave frequency. These length changes produce opposite phase shifts between two light beams in interferometer arms, when interference occurs at the beam splitter of Michelson interferometer. The resulting phase shift of a single beam of light spending time $`\tau `$ in the interferometer can be written as meers $$\delta \varphi =h\frac{\omega }{\mathrm{\Omega }}\mathrm{sin}\frac{\mathrm{\Omega }\tau }{2},$$ (1) where $`\omega `$ is the light frequency. This phase shift results an intensity signal change of the light from interferometer beam splitter hitting the photodetector. The main problem of the acoustic and interferometric methods that they both deal with gravitational waves with extremely small amplitudes of the order $`h10^{21}`$ schutz reached the Earth from deep space. One can see from eq. (1) that for gravity wave frequencies in the $`1`$ kHz range, $`\mathrm{\Omega }10^3`$ Hz, and for the light in visual frequency range, $`\omega 10^{14}`$ Hz, one has the maximum phase shift of the order $`\delta \varphi 10^{10}`$ for interferometer arms length of the order $`150`$ kilometers. Such extraordinarily weak effect requires a exceedingly high detector sensitivity both acoustic and interferometric detectors. Alternatively, GW detection may be based on effects associated with propagation of light or electromagnetic waves in gravitational fields. There are two primary effects for the light in constant gravitational field i) the deflection of light rays near massive bodies landau\_2 and ii) the Shapiro effect accounting for integrated time delay of the signal passing near a strong source of gravitational field shapiro . The same effects have to be observed for light propagating in gravity waves: the gravitational waves have to induce a weak time dependent deflection of light ray propagating through these waves and also lead to gravity-wave-induced variation in time delay. The idea to use astrometry to detect periodic variation in apparent angular separation of appropriate light sources was proposed by Fakir fakir\_1 . It is shown that for a gravity wave source located between the Earth and the light source (with line of sight close aligned to gravity wave source) a periodical variation of the order $`\mathrm{\Delta }\varphi \pi h(\mathrm{\Lambda })`$ in the angular position of the light source has to be observed. Here $`h(\mathrm{\Lambda })`$ is the dimensionless strength of the gravity wave at distances of the order gravitational wavelength $`\mathrm{\Lambda }`$ which is many orders of magnitude greater than the strength of the same waves when they reach the Earth. On the other hand one can directly measure the variation of the integrated time delay induced by gravity waves on the light emitted by distant star which passing through space region with strong gravity wave. The idea to use timing observation for detection of the gravitational waves was suggested first by Sazhin sazhin\_1 and then this problem was studied in details by several authors fakir\_2 ; damour ; kopeikin . The estimations carried out in Ref. fakir\_2 give the following answer for the rate of change in the gravity-wave-induced time delay, $`\dot{\tau }`$: $$|\dot{\tau }||h(r=D)|$$ (2) where $`h(r=D)`$ is the gravity wave strength at distances of the order of impact parameter $`D`$ for the light beam passing near the gravity wave source. To put some numbers let us consider a dissymmetric rotating neutron star with spin frequency of the order $`10^3`$ Hz and gravity wave amplitude of the order $`H10^5`$ cm thorne , then eq. (2) gives $`|\dot{\tau }|10^{12}`$ while the same effect measured near Earth results $`|\dot{\tau }|10^{26}`$ assuming that the neutron star to be at a typical distance of a few kiloparsecs from the Earth. One can see that all effects due to gravitational waves near the gravity wave source are several orders of magnitude stronger than the same ones on the Earth. In the present paper we want to suggest a new gravitational waves detection method based on the interaction of the photon with gravitational waves. Assuming that the photons from distant star passing near gravitational wave source, where the photon-gravity wave interaction assumed to be strong, the photon-gravity wave interaction leads to relatively strong modulation in time of photon frequency. The latter allows to analyze this modulation of the photons reaching the Earth by means of standard optical methods including the Fabry-Perot analysis and quantum photon correlations measurements. It is important that while the interaction of photons with gravitational wave is rather weak (proportional to strength of gravitational wave) the frequency modulation can be accumulated over large distances during the photon propagation that could result in an experimentally measurable effect on the Earth. In some aspects our treatment of the photon-gravity wave interaction resembles the effect of photon acceleration by gravitational wave mendonca , where the photon propagating in plane gravitational wave long enough time acquires a considerable increase of the frequency. However while in Ref. mendonca the photon-gravity wave interaction was treated in the frame of reference of the photon, we have considered this interaction in the point of observation frame coordinate which seems to give rather the widening of the initial monochromatic photon wave packet than the increase of the photon frequency. ## II Propagation of light near the localized source of gravitational waves. In this section we consider how the plane electromagnetic waves interact with gravitational wave field emitted by some localized source. The situation we have in mind is depicted in Fig. 1. The light signal originates from a distant star $`S`$ and travels towards the Earth to be detected by an observer $`O`$, but along the way it interacts with a gravitational body $`M`$, which also emits gravity waves. We consider the trajectory of the light ray in the two dimensional plane formed by the star $`S`$, the body which emits the GW, $`M`$, and the Earth. Provided that the wave length of the GW is large compared to the wave length of the light signal, we will use the eikonal or geometrical optics approximation to describe the interaction of the light in gravitational waves. The propagation of the electromagnetic waves in the field of gravitational waves was considered many times and we can refer the reader to several papers dealing with this questions in the geometrical optics approximation cooperstock ; estabrook ; forward ; braginsky ; lobo ; sazhin\_2 . Our subsequent analysis of this problem will be carried out in close analogy with the paper of Sazhin and Maslova sazhin\_2 , where they have considered the structure of electromagnetic field in Fabry-Perot resonator in the field of gravitational waves. However in the present paper we will be interested in a different aspect of the interaction of the light beam passing in immediate proximity to the gravitational wave source, where the amplitude of the gravitational waves is assumed to be strong in comparison with the amplitude of the same waves reaching the Earth. It is important that since the light interacts with gravitational waves in the region where they are strong, the resulting photons frequency modulation seems to be appreciable to detect it on the Earth. ### II.1 The eikonal equation. In the geometrical optics approximation the propagation of the light ray is described by the eikonal equation: $$g^{ik}\frac{\psi }{x^i}\frac{\psi }{x^k}=0,$$ (3) where $`g^{ik}=g_0^{ik}h^{ik}`$ the metric tensor associated with the Schwarzschildian static metric $`g_0^{ik}`$ of the object $`M`$, which is perturbed by the tensor $`h^{ik}`$ associated with the GW emitted from $`M`$ landau\_2 . Here we will neglect the true form of the static metric $`g_0^{ik}`$ assuming that light ray propagates in the flat Minkovsky space $`g_0^{ik}=\eta ^{ik}=\text{diag}\{1,1,1,1\}`$. It is known that the static metric leads both to the deflection of the light ray near the massive body and time delay of the light signal passing along the ray (so called Shapiro effect). It can be rigorously shown in the subsequent analysis that the former effect of the light deviation leads to the negligible correction to the additional phase accumulated by photons due to the interaction with GW and one can assume the light trajectory to be a straight line. As for static Shapiro time delay it can be easily incorporated to the final answer for photon phase and actually does not affect on the our subsequent detection method of modulated photons. Let $`\psi _0=\omega t+ky`$ be the eikonal in the absence of gravitational waves perturbation with metric $`\eta ^{ik}`$ (we assumed the light ray to be propagating along the $`y`$ axis in the plane formed by three bodies). Assuming that the perturbation is small, in the presence of gravitational waves the eikonal becomes $`\psi =\psi _0+\psi _1`$, where $`\psi _1`$ a small addition to the eikonal $`\psi _0`$ computed to first order in $`h^{ik}`$ \- it satisfies the equation: $$\frac{\psi _1}{t}+c\frac{\psi _1}{y}=\frac{\omega }{2}F(t,r),$$ (4) where we have introduced the notation $`F(t,r)=h^{00}+h^{yy}2h^{0y}`$. Using the Green’s function formalism one can find the general solution of this equation. Then to first order in the gravitational wave perturbation, $`h^{ik}`$, the additional phase acquired by the photon due to the interaction with the GW, can be written for arbitrary time dependence of the perturbed metric: $$\psi _1(t,y)=\frac{1}{2}\frac{\omega }{c}\underset{\mathrm{}}{\overset{y}{}}F(t\frac{yy^{}}{c},y^{})𝑑y^{}$$ (5) ### II.2 The emission of GW. Assuming that the GW tensor, $`h^{ik}`$, is a small perturbation to the static flat metric in the quadrupole approximation one can write the coordinate components of the GW tensor as landau\_2 : $$\stackrel{~}{h}_{\alpha \beta }=\frac{2G}{c^4}\frac{\ddot{Q}_{\alpha \beta }(tr/c)}{r}$$ (6) where $`G`$ is the Newtonian gravitation constant, $`\stackrel{~}{h}_{ik}=h_{ik}\frac{1}{2}\eta _{ik}h`$ (where $`hh_i^i`$), $`Q_{\alpha \beta }`$ is the quadrupole moment: $$Q_{\alpha \beta }(t)=\stackrel{~}{T}_{00}(t,r)x^\alpha x^\beta d^3x,$$ (7) where $`\stackrel{~}{T}_{ik}=T_{ik}\frac{1}{2}\eta _{ik}T`$, $`T_{ik}`$ is the energy-momentum tensor of the GW source, $`TT_i^i`$. Eq. (6) implies that the so called the harmonic gauge condition, $`_k\stackrel{~}{h}_i^k=0`$ has been chosen for the components of GW tensor $`h_{ik}`$. Using the gauge condition one can find the $`\stackrel{~}{h}_{00}`$ and $`\stackrel{~}{h}_{0\alpha }`$ components of the GW tensor. To do this let us write the equation $`_k\stackrel{~}{h}_i^k=0`$ separately for coordinate and time components: $$\frac{\stackrel{~}{h}_{\alpha \beta }}{x^\beta }=\frac{\stackrel{~}{h}_{\alpha 0}}{x^0},\frac{\stackrel{~}{h}_{0\alpha }}{x^\alpha }=\frac{\stackrel{~}{h}_{00}}{x^0}.$$ (8) Combining the eq. (6) with the first eq. (8) the $`\stackrel{~}{h}^{\alpha 0}`$ components of GW tensor can be written as: $$\stackrel{~}{h}_{\alpha 0}=\frac{2G}{c^3}\frac{}{x^\beta }\left[\frac{\dot{Q}_{\alpha \beta }(tr/c)}{r}\right],$$ (9) while from this equation and the second equation (8) one has: $$\stackrel{~}{h}_{00}=\frac{2G}{c^2}\frac{^2}{x^\alpha x^\beta }\left[\frac{Q_{\alpha \beta }(tr/c)}{r}\right].$$ (10) Using the expressions (6), (9),(10) for the components of the GW tensor write the explicit expression for $`F(t,r)=h^{00}+h^{yy}2h^{0y}`$ in the r. h. s. of eq. (5). Taking coordinate derivatives and keeping in mind that along the ray trajectory $`_xr=D/r`$, $`_yr=y/r`$, $`_zr=0`$ (where $`D`$ is the impact parameter) the $`F(t,r)`$ takes the form $`F(t,r)`$ $`=`$ $`{\displaystyle \frac{2G}{c^4}}\left[{\displaystyle \frac{\ddot{Q}_{xx}}{r}}{\displaystyle \frac{D^2}{r^2}}+2{\displaystyle \frac{\ddot{Q}_{xy}}{r}}{\displaystyle \frac{D}{r}}\left({\displaystyle \frac{y}{r}}1\right)+{\displaystyle \frac{\ddot{Q}_{yy}}{r}}\left({\displaystyle \frac{y}{r}}1\right)^2\right]`$ (11) $`+`$ $`{\displaystyle \frac{2G}{c^3}}\left[3{\displaystyle \frac{\dot{Q}_{xx}}{r^2}}{\displaystyle \frac{D^2}{r^2}}+2{\displaystyle \frac{\dot{Q}_{xy}}{r^2}}{\displaystyle \frac{D}{r}}\left(3{\displaystyle \frac{y}{r}}1\right)+{\displaystyle \frac{\dot{Q}_{yy}}{r^2}}\left(3{\displaystyle \frac{y^2}{r^2}}2{\displaystyle \frac{y}{r}}1\right)\right]`$ $`+`$ $`{\displaystyle \frac{2G}{c^2}}\left[3{\displaystyle \frac{Q_{xx}}{r^3}}{\displaystyle \frac{D^2}{r^2}}+6{\displaystyle \frac{Q_{xy}}{r^3}}{\displaystyle \frac{yD}{r^2}}+{\displaystyle \frac{Q_{yy}}{r^3}}\left(3{\displaystyle \frac{y^2}{r^2}}1\right)\right],`$ where all quadrupole moments $`Q_{\alpha \beta }`$ are assumed to be the functions of the retarded variable $`tr/c`$. ### II.3 Solution of the eikonal equation. For simplicity the source of GW is assumed to emit only one frequency $`\mathrm{\Omega }`$. We consider the two following situations: either the GW source has been emitting since $`t=\mathrm{}`$, or the signal is bounded in time, with an exponential decay. In the former case one can write $`\ddot{Q}_{\alpha \beta }(t)=q_{\alpha \beta }\mathrm{sin}\mathrm{\Omega }t`$. Then substituting the expression for $`F(t,r)`$, see eq. (11), into eq. (5) one could write the solution for $`\psi _1`$ as an integral over the variable $`z=y^{}/D\sqrt{1+(y^{}/D)^2}`$: $`\psi _1(\xi )`$ $`=`$ $`2{\displaystyle \frac{\omega }{c}}{\displaystyle \underset{D/(2y)}{\overset{\mathrm{}}{}}}{\displaystyle \frac{z^3H_{yy}2z^2H_{xy}+zH_{xx}}{(z^2+1)^2}}\mathrm{sin}\left[\mathrm{\Omega }\xi 2\pi {\displaystyle \frac{D}{\mathrm{\Lambda }}}z\right]𝑑z`$ (12) $`+`$ $`4{\displaystyle \frac{\omega }{c}}\left({\displaystyle \frac{\mathrm{\Lambda }}{2\pi D}}\right){\displaystyle \underset{D/(2y)}{\overset{\mathrm{}}{}}}{\displaystyle \frac{z^2(z^22)H_{yy}2(2z^3z)H_{xy}+3z^2H_{xx}}{(z^2+1)^3}}\mathrm{cos}\left[\mathrm{\Omega }\xi 2\pi {\displaystyle \frac{D}{\mathrm{\Lambda }}}z\right]𝑑z`$ $`+`$ $`4{\displaystyle \frac{\omega }{c}}\left({\displaystyle \frac{\mathrm{\Lambda }}{2\pi D}}\right)^2{\displaystyle \underset{D/(2y)}{\overset{\mathrm{}}{}}}{\displaystyle \frac{z(z^44z^2+1)H_{yy}6z^2(z^21)H_{xy}+6z^3H_{xx}}{(z^2+1)^4}}\mathrm{sin}\left[\mathrm{\Omega }\xi 2\pi {\displaystyle \frac{D}{\mathrm{\Lambda }}}z\right]𝑑z.`$ where the variable $`\xi =ty/c`$ accounts for the effect of retardation of the light signal during its propagation to the point of observation, $`H_{\alpha \beta }=2Gq_{\alpha \beta }/c^4`$, $`\mathrm{\Lambda }`$ is the wavelength of gravitational waves. Assuming that at the point of observation $`yD`$, one can safely set the lower limit for all integrals in this expression equal to zero. Then for large values of the impact parameter $`2\pi D/\mathrm{\Lambda }1`$ one can asymptotically estimate these integrals and the main contribution to $`\psi _1`$ is given by the following expression: $`\psi _1(\xi )`$ $``$ $`{\displaystyle \frac{1}{2\pi ^2}}\left({\displaystyle \frac{\mathrm{\Lambda }}{D}}\right)^2{\displaystyle \frac{\omega }{c}}H_{xx}\mathrm{sin}\mathrm{\Omega }\xi `$ (13) $``$ $`{\displaystyle \frac{h_{xx}}{\pi }}{\displaystyle \frac{\mathrm{\Lambda }}{\lambda }}\left({\displaystyle \frac{\mathrm{\Lambda }}{D}}\right)^2\mathrm{sin}\mathrm{\Omega }\xi ,`$ (14) where in the last line the answer is rewritten in terms of the dimensionless strength of gravitational waves: $`h_{xx}=H_{xx}\mathrm{\Lambda }^1`$. One could see that the main contribution to the photon phase $`\psi `$ comes from the $`h_{xx}`$ component of gravitational radiation confirming the well known fact damour ; kopeikin that only gravitational waves propagating perpendicular to the photon wave vector (directed along $`y`$ axis in our geometry) have a considerable effect on the electromagnetic radiation. Let us now consider the case where gravity waves are emitted starting at $`t=0`$ decaying exponentially with characteristic time $`\tau _0`$. In this case let us choose $`\ddot{Q}_{\alpha \beta }(t)`$ to be equal to: $$\ddot{Q}_{\alpha \beta }(t)=q_{\alpha \beta }\theta (t)\mathrm{exp}\left(\frac{t}{\tau _0}\right)\mathrm{sin}\mathrm{\Omega }t,$$ (15) where $`\theta (x)=0`$ for $`x<0`$ and $`1`$ for $`x>0`$. Then for large values of the impact parameter, $`2\pi D/\mathrm{\Lambda }1`$, the solution for $`\psi _1`$ is given by the following formula: $`\psi _1(\xi )`$ $``$ $`{\displaystyle \frac{h_{xx}}{\pi }}\left({\displaystyle \frac{\mathrm{\Lambda }}{D}}\right)^2{\displaystyle \frac{\mathrm{\Lambda }}{\lambda }}\theta (\xi )e^{\xi /\tau _0}\mathrm{sin}\mathrm{\Omega }\xi .`$ (16) One can see that in this case at the point of observation the modulated contribution to the eikonal also decaying exponentially with the same characteristic time $`\tau _0`$. ### II.4 The pre-exponential factor. Let us now go beyond the geometrical optics or eikonal approximation and calculate the pre-exponential factor for the photon wave function. The photon wave function $`\phi (t,y)`$ with momentum $`k`$ satisfies the wave equation: $$\frac{1}{\sqrt{g}}\frac{}{x^i}\left(\sqrt{g}g^{ik}\frac{\phi }{x^k}\right)=0,$$ (17) where $`g=detg_{ik}`$. Up to the first order contribution in $`h^{ik}`$ one could find $`g=(1+h)`$ with $`h=h_i^i`$. Taking the derivative over $`x^i`$ and keeping only the first order terms in $`h^{ik}`$ one could arrive to expression: $$g^{ik}\frac{^2\phi }{x^ix^k}\frac{\phi }{x^k}\frac{}{x^i}\left(h^{ik}\frac{1}{2}\eta ^{ik}h\right)=0.$$ (18) One could see that under the harmonic gauge condition $`_i\stackrel{~}{h}_k^i=0`$ with $`\stackrel{~}{h}_k^i=h_k^i\frac{1}{2}\eta _k^ih`$ the second term in above equation vanishes and the photon wave equation reduces to $$g^{ik}\frac{^2\phi }{x^ix^k}=0.$$ (19) Let us write the photon wave function as $`\phi (t,y)=c(t,y)e^{i\psi }`$, where $`\psi `$ is the eikonal of the photon interacting with gravitational waves. Then assuming that $`c=1+a`$, where $`a`$ is the small contribution of the first order in $`h^{ik}`$, one could rewrite the above equation as $$g^{ik}\left(\frac{^2a}{x^ix^k}+2i\frac{\psi }{x^i}\frac{a}{x^k}+i(1+a)\frac{^2\psi }{x^ix^k}(1+a)\frac{\psi }{x^i}\frac{\psi }{x^k}\right)=0.$$ (20) The last term in this equation vanishes due to the eikonal equation (3). Since the free eikonal $`\psi _0=ky\omega t`$ is the linear function of $`x^i`$ the all second derivatives of $`\psi _0`$ are vanish and the remaining wave equation up to the first order terms in $`h^{ik}`$ takes the form: $$\eta ^{ik}\frac{^2a}{x^ix^k}+2i\eta ^{ik}\frac{\psi _0}{x^i}\frac{a}{x^k}+i\eta ^{ik}\frac{^2\psi _1}{x^ix^k}=0.$$ (21) Assuming that the $`a(t,y)=a(ty/c)`$ thus the first term above vanishes one could finally find $$\frac{a}{t}+c\frac{a}{y}=\frac{1}{2}\frac{c^2}{\omega }\left(\frac{1}{c^2}\frac{^2\psi _1}{t^2}\frac{^2\psi _1}{y^2}\right).$$ (22) Substituting here the general solution for $`\psi _1`$ (see eq. (5)) one can express the r. h. s. of eq. (22) through the function $`F(t,y)`$: $$\frac{a}{t}+c\frac{a}{y}=\frac{c}{4}\left(\frac{F}{y}\frac{1}{c}\frac{F}{t}\right)$$ (23) According to our previous analysis the main contribution to $`a(t,y)`$ is given by the $`H_{xx}`$ component of gravitational radiation: $$F(t,r)H_{xx}\frac{\mathrm{sin}\mathrm{\Omega }(tr/c)}{r}\frac{D^2}{r^2}$$ (24) Substituting this expression to the eq. (23) the solution for $`a(\xi )`$ for large values of impact parameter $`2\pi D/\mathrm{\Lambda }1`$ has the form: $$a(\xi )\frac{h_{xx}}{\pi }\left(\frac{\mathrm{\Lambda }}{D}\right)^2\mathrm{cos}\mathrm{\Omega }\xi $$ (25) Combining together eq. (14) with eq. (25) one can finally write the photon wave function at the point of observation in the form: $$\phi _k(\xi )=\frac{e^{i\omega (\xi +\tau _g\mathrm{sin}\mathrm{\Omega }\xi )}}{1+\tau _g\mathrm{\Omega }\mathrm{cos}\mathrm{\Omega }\xi },$$ (26) where we have introduced the time $`\tau _g`$ associated with integrated Shapiro time delay (see below): $$\tau _g=\frac{h_{xx}}{\pi }\left(\frac{\mathrm{\Lambda }}{D}\right)^2\mathrm{\Omega }^1.$$ (27) The same solution eq. (26) for the photon wave function is valid for the case where gravitational radiation are bounded exponentially in time (see eq. (15)) however in this case $`\tau _g`$ depends on time and equals to $$\tau _g(\xi )=\frac{h_{xx}}{\pi }\left(\frac{\mathrm{\Lambda }}{D}\right)^2\mathrm{\Omega }^1\theta (\xi )e^{\xi /\tau _0}.$$ (28) ### II.5 Modulation due to alternating Newtonian gravitational field. The observation and detection of the gravitational waves undoubtedly provides us a new astrophysical tool which can give us a deeper insight on the processes like the neutron star formation, the processes happening in close binaries at the final stage of it evolution, etc…. However the observation of gravitational waves gives us also an additional proof of the General Theory of Relativity itself and it seems reasonable to consider the effect of photon phase modulation in Newtonian theory of gravitation where the gravitational potential instantly adjusts to the current configuration of moving bodies. Consider for example an oscillating neutron star or a rotating double star system which in Newtonian theory of gravity causes an alternating Newtonian gravitational potential $`\varphi (𝐫,t)`$. This alternating field can also lead to photon phase modulation and results in principle in the same effect like for the case of photon in the field of gravitational waves. The aim of this section is to compare the effect of photon phase modulation between these two theories. As the rotation or the oscillation frequency of the stars is much less than the frequency of the photon we can work in the same eikonal approximation as in the previous sections. The Newtonian gravitational field contributes only to the diagonal metric elements: $`h^{00}=h^{ii}=2\varphi /c^2`$, where $`\varphi `$ \- is a gravitational potential at the point of observation. At large distances compared to the size of the star system we can leave only quadrupole term in the potential $$\varphi (𝐫,t)=GQ_{\alpha \beta }\frac{n_\alpha n_\beta }{2r^3}.$$ (29) Here $`r`$ \- is the distance from the star to the photon, the coordinate axis are chosen to be the same as before, so $`r=\sqrt{y^2+D^2}`$, $`n_i`$ \- the direction vectors of $`𝐫`$: $`n_x=D/\sqrt{y^2+D^2}`$, $`n_y=y/\sqrt{D^2+y^2}`$, $`n_z=0`$, $`Q_{\alpha \beta }(t)`$ is the alternating tensor of the mass quadrupole moment of the star. The values of $`Q_{\alpha \beta }`$ are determined by the parameters of the star, like orientation of the angular velocity, eccentricity, etc. We will further discuss the general case, assuming only, that due to rotation or oscillation $`\ddot{Q}_{\alpha \beta }(t)=q_{\alpha \beta }\mathrm{sin}\mathrm{\Omega }t`$. The correction to the eikonal due to rotation is given by the same equation (5) as for the GW case solution of the eikonal equation, where in our case $`F(t,y)=h^{00}+h^{yy}=2h^{00}`$ $$\mathrm{\Psi }_n=\frac{\omega }{c}\underset{\mathrm{}}{\overset{y}{}}h^{00}(t\frac{yy^{}}{c},y^{})𝑑y^{}$$ (30) Substituting the expression for $`\varphi `$ into the formula above and making the substitution $`z=y^{}/D`$ one obtains: $`\mathrm{\Psi }_n(\xi )`$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle \frac{\omega }{c}}\left({\displaystyle \frac{\mathrm{\Lambda }}{2\pi D}}\right)^2{\displaystyle \underset{\mathrm{}}{\overset{y/D}{}}}{\displaystyle \frac{H_{xx}+2H_{xy}z+H_{yy}z^2}{(1+z^2)^{5/2}}}\mathrm{sin}\left[\mathrm{\Omega }\xi +2\pi {\displaystyle \frac{D}{\mathrm{\Lambda }}}z\right]𝑑z,`$ (31) where we have introduced $`H_{\alpha \beta }=2Gq_{\alpha \beta }/c^4`$. Then for large values of the impact parameter $`2\pi D/\mathrm{\Lambda }1`$ and large distances $`y/D1`$ one can obtain the following approximate expression for $`\mathrm{\Psi }_n`$: $`\mathrm{\Psi }_n(\xi )={\displaystyle \frac{1}{6}}{\displaystyle \frac{\omega }{c}}\sqrt{{\displaystyle \frac{\mathrm{\Lambda }}{D}}}e^{2\pi D/\mathrm{\Lambda }}\left[(H_{xx}H_{yy})\mathrm{sin}\mathrm{\Omega }\xi +2H_{xy}\mathrm{cos}\mathrm{\Omega }\xi \right].`$ (32) One can see that the photon phase modulation due to the Newtonian alternating potential vanishes exponentially for large impact parameters in contrast to the GW modulation which vanishes as $`1/D^2`$ (see eq. (14)). Technically the difference between Newtonian and GW cases follows from the different $`y^{}`$ dependence of the $`F(t,y^{})`$ function. For the properly retarded GW radiation the solution for $`\psi _1`$, see eq. (14), involves the oscillating integral over $`\mathrm{sin}\mathrm{\Omega }[\xi +(y^{}\sqrt{(y^{})^2+D^2})/c]`$, where the oscillating dependence on $`y^{}`$ near the source $`|y^{}|D`$ almost vanishes making the light-GW interaction more effective. Oppositely for the instantly adjusted Newtonian radiation one has in eq. (30) fast oscillating integral over $`y^{}`$ with $`\mathrm{sin}\mathrm{\Omega }(\xi +y^{}/c)`$ near vicinity of the gravitational wave source, which leads to a very small value of modulation. ### II.6 The Shapiro effect in GW. Using the results of the previous sections let us calculate the integrated time delay of the light signal propagating in the field of gravitational waves. Consider an arbitrary wave packet emitted by some distant light source traveling toward to Earth near a source of GW: $$\phi (\xi )=\phi _\omega e^{i\omega \xi }\frac{d\omega }{2\pi }$$ (33) where $`\xi =ty/c`$. Using the results for monochromatic, photons propagating in the field of gravitational radiation one can find the signal at the point of observation: $`\phi _{ob}(\xi )`$ $`=`$ $`{\displaystyle \phi _\omega e^{i\omega [\xi +\tau _g\mathrm{sin}\mathrm{\Omega }\xi ]}\frac{d\omega }{2\pi }}`$ (34) $`=`$ $`\phi (\xi +\tau _g\mathrm{sin}\mathrm{\Omega }\xi ).`$ One can see that the interaction with the gravitational radiation leads to alternating time delay of the signal $`\delta t=\tau _g\mathrm{sin}\mathrm{\Omega }t`$, (note that the $`\tau _g`$ is negative, see eq. (27)). This effect was suggested independently first by Sazhin and Detweiler sazhin\_1 for detection of gravitational waves using the timing observation of a pulsar the line of sight to which passes near the source of GW. ## III Detection of the GW Let us now turn to the discussion about how the photons modulated by the interaction with the gravitational waves can be measured in some realistic setup. Consider first the simple case where the modulated photons (26) simply hit a photodetector which can react at all photon’s frequencies. However, any real photodetector reacts not on the vector potential $`𝐀(t)`$ of the photons in the photodetector directly, but rather on the electrical field, $`𝐄(t)`$, induced by the photons in this photodetector. Then the one photon photo-detection probability, $`P(t)\mathrm{\Delta }t`$, to observe a photon during time interval $`\mathrm{\Delta }t`$ can be expressed as mandel : $$P(t)\mathrm{\Delta }t=\alpha \frac{Sc\mathrm{\Delta }t}{E_{det}}\widehat{𝐄}^{()}(t,𝐫)\widehat{𝐄}^{(+)}(t,𝐫)$$ (35) where $`S`$ is the area of the photodetector, $`E_{dec}`$ some characteristic energy describing the interaction of the photodetector with incoming photons, $`\alpha `$ is the quantum efficiency of the photo-detection, $`\widehat{𝐄}^{(+)}(t,𝐫)`$ is the positive frequency part of the electrical field operator, $`\widehat{𝐄}^{()}(t,𝐫)`$ is the Hermitian conjugate of $`\widehat{𝐄}^{(+)}(t,𝐫)`$: $$\widehat{𝐄}^{(+)}(t,𝐫)=\frac{\sqrt{4\pi }}{L^{3/2}}\underset{𝐤s}{}\sqrt{\frac{\mathrm{}}{2\omega _k}}\dot{\phi }_k(t,𝐫)𝐞_{ks}\widehat{a}_{ks}.$$ (36) At the point of observation we set $`𝐫=0`$ to shorten the notations. Taking the time derivative of the photon wave function (26) one gets the photo-detection probability equals to $`P(t)\mathrm{\Delta }t`$ $`=`$ $`{\displaystyle \frac{4\pi \alpha Sc\mathrm{\Delta }t}{E_{det}L^3}}{\displaystyle \underset{𝐤s}{}}{\displaystyle \frac{\mathrm{}\omega _k}{2}}n_{ks}`$ (37) $`=`$ $`2\alpha {\displaystyle \frac{S}{r^2}}\mathrm{\Delta }t{\displaystyle 𝑑\omega \frac{\mathrm{}\omega }{E_{det}}n(\omega )},`$ where $`r`$ is the distance between the light source and the Earth, $`n(\omega )d\omega `$ is the number of the photons with frequency $`\omega `$, emitted by the distant light source per unit time, into a unit angular domain. One can see that the photo-detection probability does not feel any photon phase modulation due to GW and this probability is the same as it would be for the monochromatic photons emitted by a distant light source. Thus for such a simple setup it is not possible to retrieve any information about the gravity wave from the photodetector signal. Consider however a more realistic setup where the photodetector has a sensitivity edge $`\omega _s`$ which means that photodetector does not react on the photons with frequency less than $`\omega _s`$. One can treat this situation as if the photons passed through a filter with transparency $`t(\omega )=\theta (\omega \omega _s)`$, before hitting the photodetector. It is convenient to write this transparency as an operator acting on the photon wave function: $`\widehat{t}=\theta [i_t\omega _s]`$. Then after passing such a filter the photon has a wave function equals to $`\widehat{t}\phi _k(t)\theta [\omega _k(t)\omega _s]\phi _k(t)`$, where $`\omega _k(t)=\omega _k[1+\mathrm{\Omega }\tau _g\mathrm{cos}\mathrm{\Omega }t]=\omega _kf(t)`$. Then for the photo-detection probability one has: $`P(t)\mathrm{\Delta }t2\alpha {\displaystyle \frac{S\mathrm{\Delta }t}{r^2}}{\displaystyle 𝑑\omega \frac{\mathrm{}\omega }{E_{det}}\theta [\omega f(t)\omega _s]n(\omega )}.`$ (38) Consider now the case where photons coming to the photodetector from a sharp Lorentzian spectral line with maximum, $`\omega _0`$, near the photodetector sensitivity edge $`\omega _0\omega _s`$ and width $`\mathrm{\Gamma }_0`$: $$n(\omega )=n(\omega _0)\frac{\mathrm{\Gamma }_0^2}{(\omega \omega _0)^2+\mathrm{\Gamma }_0^2},$$ (39) then according to the eq. (38) the photo-detection probability for the modulated photons (26) has an appreciable contribution which is periodic in time at the GW frequency, and which is proportional to the strength of gravitational waves: $$P(t)\mathrm{\Delta }tP_0\mathrm{\Delta }t\left(1+\mathrm{\Omega }\tau _g\frac{2\omega _0}{\pi \mathrm{\Gamma }_0}\mathrm{cos}\mathrm{\Omega }t\right),$$ (40) where $`P_0\mathrm{\Delta }t`$ is the photo-detection probability for non modulated photons. From the above analysis one can conclude that in order to detect the modulation of the photons by GW one should place before the photodetector some filter which has a finite frequency band. Then the formula for photo-detection probability (38) with a sensitivity edge has a rather general sense. In fact let us consider an arbitrary filter placed before the photodetector with transparency $`T(\omega )`$. Then provided that the time spent by the photon in this filter (that is an inverse frequency band of the filter) is much smaller than frequency of the GW, $`\mathrm{\Omega }`$, one can treat the modulated photons as monochromatic photons with slowly varying frequency $`\omega _k(t)=\omega _k[1+\mathrm{\Omega }\tau _g\mathrm{cos}\mathrm{\Omega }t]=\omega _kf(t)`$. In this case the photo-detection probability can be written simply as $$P(t)\mathrm{\Delta }t2\alpha \frac{S\mathrm{\Delta }t}{r^2}𝑑\omega \frac{\mathrm{}\omega }{E_{det}}T(\omega f(t))n(\omega ).$$ (41) From this expression one can see that there are two different regimes of GW detection in this proposed setup. Consider the first regime where the distribution function of the light source, $`n(\omega )`$, has a sharp peak of the width $`\mathrm{\Gamma }_0`$ (or in more general case some sufficient irregularity like the edge of the spectrum) centered near the transmission window of the filter. Then, provided that the amplitude of the photon’s frequency modulation, $`\omega _k\mathrm{\Omega }\tau _g`$, is much bigger than the width of the spectrum irregularity, $`\omega _k\mathrm{\Omega }\tau _g\mathrm{\Gamma }_0`$, the resulting photo-detection signal will be a periodic sequence of peaks. In the opposite case where the light spectrum $`n(\omega )`$ is a slowly varying function of the frequency within the transparency window of the filter one can consider $`n(\omega )`$ as a constant in eq. (41). Then the time dependence of the photo-detection probability can be written as: $$P(t)\mathrm{\Delta }t2\alpha \frac{S\mathrm{\Delta }t}{r^2}\frac{1}{f(t)}𝑑\omega \frac{\mathrm{}\omega }{E_{det}}T(\omega )n(\omega ).$$ (42) This result can be understood as the time modulation of the number of incoming photons per unit time within a finite frequency interval. From another point of view one can note that the modulated photon wave function (26) can be written as a superposition of waves with energies shifted by $`n\mathrm{\Omega }`$ ($`n`$ integer): $$\phi _k(t)=\frac{1}{1+\mathrm{\Omega }\tau _g\mathrm{cos}\mathrm{\Omega }t}\underset{\mathrm{}}{\overset{+\mathrm{}}{}}J_n(\omega \tau _g)e^{i(\omega +n\mathrm{\Omega })t}$$ (43) where $`J_n`$ is the Bessel function. In order to detect a gravity wave signal it is therefore necessary to provoke an interference between these Fourier components. To do it one can both investigate the modulated light signal passing through the Fabry-Perot interferometer or from the other hand study the intensity-intensity correlation of the photons coming on the photodetector. ### III.1 Analysis of the light signal with an interferometer In this section we consider in detail the setup where at the point of observation - the Earth - the modulated light signal before hitting the photodetector passes through an interferometer, which is characterized by a complex transmission amplitude $`t(\omega )`$. We will assume that the light passing through a Fabry-Perot filter comes from a single Lorentzian spectral line of the type (39). For simplicity, let us first assume that there is only one FP resonance within this spectral line. The transparency of FP is written in the usual way: $$t(\omega )=\frac{i\mathrm{\Gamma }/2}{\omega \omega _0+i\mathrm{\Gamma }/2}.$$ (44) Here a derivation of the photo-detection probability is provided. It relies on the assumption $`\mathrm{\Omega }\tau _g1`$, which is relevant for experimental situations (see below). Consider the propagation of a photon wave-packet $`\phi _k(t)`$ (26) through the FP interferometer. The resulting wave packet after the FP can be written $`\varphi _k(t)=\widehat{t}\phi _k(t)`$, where $`\widehat{t}`$ is the transparency operator of the FP in real time representation: $$\widehat{t}=\frac{i\mathrm{\Gamma }}{2}[i_t\omega _0+i\mathrm{\Gamma }/2]^1,$$ (45) where $`\mathrm{\Gamma }`$ is the width of FP resonance. However the real FP filter does not act directly on the vector potential of the incoming photons but rather on the electromagnetic field induced by the these photons. Using similar arguments the positive frequency part of electrical field operator $`𝐄^{(+)}(t,r)`$, after passing the Fabry-Perot filter, can be written in terms of the transmission operator of the filter: $$\widehat{𝐄}^{(+)}(t,𝐫)=\frac{\sqrt{4\pi }}{L^{3/2}}\underset{𝐤s}{}\sqrt{\frac{\mathrm{}}{2\omega _k}}\left[\widehat{t}\dot{\phi }_k(t,𝐫)\right]𝐞_{ks}\widehat{a}_{ks}.$$ (46) With these notations the photo-detection probability can be written as $$P(t)=2\alpha \frac{S}{r^2}\frac{\mathrm{}d\omega }{\omega E_{det}}\left[\widehat{t}\dot{\phi }_k(t)\right]^{}\left[\widehat{t}\dot{\phi }_k(t)\right]n(\omega )𝑑\omega .$$ (47) The action of the operator $`\widehat{t}`$ on the function $`\dot{\phi }_k(t)`$ can be found using the Green’s function formalism: $$\widehat{t}\dot{\phi }_k(t)=\frac{\mathrm{\Gamma }}{2}\underset{\mathrm{}}{\overset{t}{}}e^{i\omega _0(t\tau )\frac{\mathrm{\Gamma }}{2}(t\tau )}\dot{\phi }_k(\tau )𝑑\tau .$$ (48) Substituting this expression into eq. (47) in the limit $`\omega _0\mathrm{\Gamma }`$ one finally obtains the following expression for the photo-detection probability: $`P(t)=2\alpha {\displaystyle \frac{S}{r^2}}{\displaystyle \frac{\mathrm{}\omega _0}{E_{det}}}{\displaystyle \frac{\mathrm{\Gamma }^2}{4}}{\displaystyle n(\omega )e^{\mathrm{\Gamma }t}𝑑\omega \underset{\mathrm{}}{\overset{t}{}}\underset{\mathrm{}}{\overset{t}{}}e^{i\omega _0(\tau s)+\frac{\mathrm{\Gamma }}{2}(\tau +s)}\dot{\phi }_k^{}(\tau )\dot{\phi }_k(s)𝑑\tau 𝑑s}`$ (49) Performing the integral over $`\omega `$ and introducing the relative time $`\xi =\tau s`$, and the total time $`\eta =(\tau +s)/2`$, in the working limit $`\mathrm{\Omega }\tau _g1`$ one can write the photo-detection probability as: $`P(t)=\alpha \pi {\displaystyle \frac{S}{r^2}}{\displaystyle \frac{\mathrm{}\omega _0}{E_{det}}}n(\omega _0)\mathrm{\Gamma }_0{\displaystyle \frac{\mathrm{\Gamma }^2}{4}}{\displaystyle \underset{\mathrm{}}{\overset{t}{}}}𝑑\eta e^{\mathrm{\Gamma }(t\eta )}{\displaystyle \underset{2(t\eta )}{\overset{2(t\eta )}{}}}e^{i\xi [\omega _0\mathrm{\Omega }\tau _g\mathrm{cos}\mathrm{\Omega }\eta ]}e^{\frac{\mathrm{\Gamma }_0(\eta )}{2}|\xi |}𝑑\xi `$ (50) Performing the integration over $`\xi `$ in the limit $`\mathrm{\Gamma }_0\mathrm{\Gamma }`$ the leading contribution to the photo-detection probability is given by the following expression: $`P(t)=\alpha \pi {\displaystyle \frac{S}{r^2}}{\displaystyle \frac{\mathrm{}\omega _0}{E_{det}}}n(\omega _0)\mathrm{\Gamma }_0{\displaystyle \frac{\mathrm{\Gamma }^2}{4}}{\displaystyle \underset{0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{\mathrm{\Gamma }_0(tz)e^{\mathrm{\Gamma }z}dz}{(\omega _0\mathrm{\Omega }\tau _g)^2\mathrm{cos}^2\mathrm{\Omega }(tz)+\mathrm{\Gamma }_0^2(tz)/4}}`$ (51) where we have made the substitution $`z=(t\eta )`$ and we introduced the notation $`\mathrm{\Gamma }_0(t)=\mathrm{\Gamma }_0|1+\mathrm{\Omega }\tau _g\mathrm{cos}\mathrm{\Omega }t|`$. #### III.1.1 ”Strong” GW or narrow spectrum. Let us consider the limit where the frequency broadening of the initially monochromatic photon wave packet due to the interaction with gravitational waves, $`\delta \mathrm{\Gamma }=\omega _0\mathrm{\Omega }\tau _g`$, is much bigger than the width of the spectral line, $`\mathrm{\Gamma }_0`$. The condition $`\delta \mathrm{\Gamma }\mathrm{\Gamma }_0`$ can be achieved both in the case where the amplitude of GW is strong enough or when one has a very narrow spectral line. In this limit one can safely neglect the dependence on $`|1+\mathrm{\Omega }\tau _g\mathrm{cos}\mathrm{\Omega }t|`$ and put $`\mathrm{\Gamma }_0(t)=\mathrm{\Gamma }_0`$ in eq. (51) assuming $`\mathrm{\Omega }\tau _g1`$. Then the leading contribution to the photo-detection probability can be written in terms of the occupation number, $`n(\omega )`$, of the spectral line: $$P(t)=\alpha \pi \frac{S}{r^2}\frac{\mathrm{}\omega _0}{E_{det}}\mathrm{\Gamma }^2\underset{0}{\overset{\mathrm{}}{}}n[\omega _0(tz)]e^{\mathrm{\Gamma }z}𝑑z$$ (52) where $`\omega _0(t)=\omega _0(1+\mathrm{\Omega }\tau _g\mathrm{cos}\mathrm{\Omega }t)`$. This result has a very simple qualitative explanation which is consistent with our previous simple arguments for the photodetector with sensitivity edge (see eq. (41)). Consider a photon wave packet coming to FP filter: $`\phi _k=𝒞(t)e^{i\omega _k(t+\tau _g\mathrm{sin}\mathrm{\Omega }t)}`$. Such a wave packet in the limit $`\mathrm{\Omega }\omega _k`$ can be interpreted as a photon with a frequency which is slowly varying in time $`\omega _k(t)=\omega _k(1+\mathrm{\Omega }\tau _g\mathrm{cos}\mathrm{\Omega }t)`$. If however the time spent by this photon in the FP filter, $`\mathrm{\Gamma }^1`$, much smaller than the characteristic time of the frequency change $`\mathrm{\Omega }^1`$ one can assume the photon at any instant of time to be monochromatic with frequency $`\omega _k(t)`$. Then one can take advantage of our eq. (41). Performing the integration over $`\omega `$ in eq. (41) with Lorentzian spectral line $`n(\omega )`$ (39) one has $$P(t)=\alpha \pi \frac{S}{r^2}\frac{\mathrm{}\omega _0}{E_{det}}n[\omega _0(t)]\mathrm{\Gamma }.$$ (53) On the other hand this result can be derived directly from our elaborate analysis by putting $`n(\omega _0(tz))=n(\omega _0(t))`$ in eq. (52). The exponential factor in eq. (52) describes a time delay of the photon wave packet propagating through the FP filter. The result eq.(53) describes the periodic sequence of symmetric peaks with a half period of gravitational waves $`\pi /\mathrm{\Omega }`$ and widths equal to $$\tau =\frac{\mathrm{\Gamma }_0}{\delta \mathrm{\Gamma }}\mathrm{\Omega }^1.$$ (54) If however the width of these peaks becomes smaller than the time needed for the photon to penetrate through the FP filter: $`\tau \mathrm{\Gamma }^1`$ then the form of peaks becomes asymmetric and the more accurate expression (52) has to be used. It should be noted that the expression (52) is valid not only for a single spectral line (39) but also for arbitrary distribution function for the light source, where there is a sufficient irregularity in $`n(\omega )`$ like, for example, at the edge of the spectrum. Indeed if one consider the arbitrary distribution function $`n(\omega )`$ which changes much at frequency scale $`\mathrm{\Delta }\omega `$ around $`\omega _0`$ then it follows from eq. (52) that one again has a periodical sequence of peaks for photo-detection signal in regime $`\delta \mathrm{\Gamma }\mathrm{\Delta }\omega `$. In this sense the described effects (52), (53) have nothing to do with the interference of the different components of the photon wave packet (43) and deal rather with the effect of the photon frequency modulation $`\omega _k(t)=\omega _k(1+\mathrm{\Omega }\tau _g\mathrm{cos}\mathrm{\Omega }t)`$ due to the interaction with the GW. If however the amplitude of the frequency modulation $`\delta \mathrm{\Gamma }`$ becomes smaller than the characteristic scale of the spectral function $`n(\omega )`$ (the case of a broad spectrum) the effects associated with frequency modulation completely disappear. ### III.2 ”Weak” GW or broad spectrum. Consider now the opposite limit to the previous case when the amplitude of GW is small or when the width of the spectral line is large enough so that the amplitude of the frequency modulation, $`\delta \mathrm{\Gamma }`$, is much smaller than the width of the spectral line, $`\mathrm{\Gamma }_0`$. It should be noted that this situation is the most common situation in practice since the amplitude of GW is extremely small even in the vicinity of GW source. In the limit $`\delta \mathrm{\Gamma }\mathrm{\Gamma }_0`$ one can safely expand the integral expression in eq. (51) up to the lowest order in the small ratio $`\delta \mathrm{\Gamma }/\mathrm{\Gamma }_01`$ and perform the integration over $`z`$. Keeping all terms of the type $`1+\mathrm{\Omega }\tau _g\mathrm{cos}\mathrm{\Omega }t`$ in the working assumption $`\mathrm{\Gamma }\mathrm{\Omega }`$ the photo-detection probability can be now written as: $`P(t)\alpha \pi {\displaystyle \frac{S}{r^2}}{\displaystyle \frac{\mathrm{}\omega _0}{E_{det}}}n(\omega _0)\mathrm{\Gamma }\left(1\mathrm{\Omega }\tau _g\mathrm{cos}\mathrm{\Omega }t2{\displaystyle \frac{\delta \mathrm{\Gamma }^2}{\mathrm{\Gamma }_0^2}}\mathrm{cos}2\mathrm{\Omega }t\right)`$ (55) One can see that in the case of a broad spectrum one obtains small oscillations against a huge constant background with frequencies $`\mathrm{\Omega }`$ and $`2\mathrm{\Omega }`$ suppressed by small factors $`\mathrm{\Omega }\tau _g1`$ and $`\delta \mathrm{\Gamma }^2/\mathrm{\Gamma }_0^21`$, correspondingly. However as one can see from eq. (55) even for the extreme case of infinitely broad spectral line still there is a time dependent contribution to the $`P(t)`$: $$P(t)\alpha \pi \frac{S}{r^2}\frac{\mathrm{}\omega _0}{E_{det}}n(\omega _0)\mathrm{\Gamma }\left[1\mathrm{\Omega }\tau _g\mathrm{cos}\mathrm{\Omega }t\right]$$ (56) It should be noted that this expression does not depend absolutely on the form of the spectra of the distant light source and the time modulation of the photo-detection signal defined only by the strength of the gravitational waves. The amplitude of the photo-detection probability in this case proportional to the number of photons $`n(\omega _0)\mathrm{\Gamma }`$ coming to the photodetector per unit time and practically one can use the FP filter with appropriate width $`\mathrm{\Gamma }`$ to collect the enough number of photons. The second important point which has to be emphasized here is that the time dependent term in eq. (56) has nothing to do with the effect of the frequency modulation mentioned above (see eq. 52, 53) and presents an absolutely different effect associated with interference of the different components of the photon’s wave packet (43) after passing the FP filter. To clarify the nature of this interference effect we will consider in the next section an example where there are two extremely narrow FP resonances, $`\mathrm{\Gamma }\mathrm{\Omega }`$, within the broad spectrum of the distant light source. ### III.3 Two resonances case. Let us consider a more general case where two narrow resonances at frequencies $`\omega _0\pm \mathrm{\Delta }/2`$ are present within the spectral line ($`\mathrm{\Delta }`$ is the separation between resonances). In this case the transmission amplitude of the FP filter, $`t(\omega )`$ again can be written as an operator acting on the e.m.f. induced by modulated photons: $`\widehat{t}=\widehat{t}_1+\widehat{t}_2`$, where $`\widehat{t}_{1(2)}`$ corresponds to the first (second) resonance: $$\widehat{t}_{1(2)}=e^{\pm i\varphi }\frac{i\mathrm{\Gamma }}{2}\left[i_t\omega _0\pm \frac{\mathrm{\Delta }}{2}+i\frac{\mathrm{\Gamma }}{2}\right]^1,$$ (57) where $`2\varphi `$ is the relative phase difference between resonances. Using similar type of arguments and the same method of calculation as for the case of FP filter with a single resonance, in the limit $`\mathrm{\Gamma }_0\mathrm{\Omega }`$, $`\mathrm{\Gamma }_0\mathrm{\Gamma }`$ one arrives at the following expression for the photo-detection probability: $$P(t)=\alpha \frac{\pi S}{r^2}\frac{\mathrm{}\omega _0}{E_{det}}n(\omega _0)\mathrm{\Gamma }_0\frac{\mathrm{\Gamma }^2}{4}\left[P_1(t)+2P_{12}(t)+P_2(t)\right]$$ (58) where $`P_{1(2)}(t)={\displaystyle \underset{0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{\mathrm{\Gamma }_0(tz)e^{\mathrm{\Gamma }z}dz}{(\delta \mathrm{\Gamma }\mathrm{cos}\mathrm{\Omega }(tz)\pm \frac{\mathrm{\Delta }}{2})^2+\frac{\mathrm{\Gamma }_0^2(tz)}{4}}}`$ (59) $`P_{12}(t)={\displaystyle \underset{0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{\mathrm{\Gamma }_0(tz)\mathrm{cos}(\mathrm{\Delta }z\varphi )e^{\mathrm{\Gamma }z}dz}{\delta \mathrm{\Gamma }^2\mathrm{cos}^2\mathrm{\Omega }(tz)+\frac{\mathrm{\Gamma }_0^2(tz)}{4}}}`$ (60) Consider now the extreme limit of infinitely broad spectrum: $`\mathrm{\Gamma }_0\mathrm{}`$, then in the limit $`\mathrm{\Gamma }\mathrm{\Omega }`$, $`\mathrm{\Gamma }\mathrm{\Delta }`$ one has: $$P(t)=\alpha \pi \frac{S}{r^2}\frac{\mathrm{}\omega _0}{E_{det}}n(\omega _0)(2\mathrm{\Gamma })[1\mathrm{\Omega }\tau _g\mathrm{cos}\mathrm{\Omega }t].$$ (61) Comparing this result with the expression for the photo-detection probability for the single resonance case (56) one can see that in the regime of a broad spectrum $`\mathrm{\Gamma }_0\mathrm{}`$ the photo-detection probability does not depend on the number of resonances in FP filter and depends rather on the total transparency of the filter ($`2\mathrm{\Gamma }`$ in present case). Consider however the opposite limit of extremely narrow resonances $`\mathrm{\Gamma }\mathrm{\Omega }`$ (this is hardly possible in practice). Then the resonance factor with respect to separation between resonances, $`\mathrm{\Delta }`$, appears in the expression for photo-detection probability: $`P(t)\alpha \pi {\displaystyle \frac{S}{r^2}}{\displaystyle \frac{\mathrm{}\omega _0}{E_{det}}}n(\omega _0)(2\mathrm{\Gamma })\left[1{\displaystyle \frac{\mathrm{\Omega }\tau _g\mathrm{\Gamma }}{2}}{\displaystyle \frac{\mathrm{\Gamma }\mathrm{cos}(\mathrm{\Omega }t\varphi )+(\mathrm{\Omega }\mathrm{\Delta })\mathrm{sin}(\mathrm{\Omega }t\varphi )}{\mathrm{\Gamma }^2+(\mathrm{\Omega }\mathrm{\Delta })^2}}\right]`$ (62) One can see that there is a parametric resonance for the photo-detection signal at $`\mathrm{\Delta }=\mathrm{\Omega }`$. The photo-detection probability at the resonance equals to: $$P_{\mathrm{\Delta }=\mathrm{\Omega }}(t)\alpha \frac{\pi S}{r^2}\frac{\mathrm{}\omega _0}{E_{det}}n(\omega _0)(2\mathrm{\Gamma })\left[1\frac{\mathrm{\Omega }\tau _g}{2}\mathrm{cos}(\mathrm{\Omega }t\varphi )\right],$$ (63) while for $`|\mathrm{\Delta }\mathrm{\Omega }|\mathrm{\Gamma }`$ (out of resonance) the term which depends on time in the photo-detection probability is saturated by the small factor $`\mathrm{\Gamma }/|\mathrm{\Delta }\mathrm{\Omega }|`$. A similar resonance appears in regime $`\mathrm{\Gamma }\mathrm{\Omega }`$ for the case where the width of the spectrum $`n(\omega )`$ is broad but $`(\omega _0/\mathrm{\Gamma }_0)^2\mathrm{\Omega }\tau _g1`$ so one can neglect all factors $`|1+\mathrm{\Omega }\tau _g\mathrm{cos}\mathrm{\Omega }t|`$ in eq. (58): $`P(t)\alpha {\displaystyle \frac{\pi S}{r^2}}{\displaystyle \frac{\mathrm{}\omega _0}{E_{det}}}n(\omega _0)(2\mathrm{\Gamma })\left[1{\displaystyle \frac{\delta \mathrm{\Gamma }^2}{\mathrm{\Gamma }_0^2}}\mathrm{\Gamma }{\displaystyle \frac{\mathrm{\Gamma }\mathrm{cos}(2\mathrm{\Omega }t\varphi )+(2\mathrm{\Omega }\mathrm{\Delta })\mathrm{sin}(2\mathrm{\Omega }t\varphi )}{\mathrm{\Gamma }^2+(2\mathrm{\Omega }\mathrm{\Delta })^2}}\right]`$ (64) One can see that the parametric resonance at $`\mathrm{\Delta }=2\mathrm{\Omega }`$ appears in this case. Now we argue that the resonances at $`\mathrm{\Delta }=\mathrm{\Omega }`$ and $`\mathrm{\Delta }=2\mathrm{\Omega }`$ (62), (64) arises because of the interference of different components of the photon’s wave packet (26). To show it let us simply write the transparency of the FP filter as $`t(\omega )=\delta _\mathrm{\Gamma }(\omega \omega _0)+\delta (\omega \omega _0\mathrm{\Delta })`$ where $`\delta _\mathrm{\Gamma }(\omega )`$ is a narrow peaked function with finite width $`\mathrm{\Gamma }\mathrm{\Omega }`$. Then using the expansion of the photon wave packet (43) one can write the photo-detection probability in terms of the sums over different components of the photon’s wave packet: $`P(t)=2\alpha {\displaystyle \frac{S}{r^2}}{\displaystyle \underset{n,m=\mathrm{}}{\overset{\mathrm{}}{}}}{\displaystyle \frac{\mathrm{}\omega }{E_{det}}n(\omega )𝑑\omega e^{i(nm)\mathrm{\Omega }t}t^{}(\omega +n\mathrm{\Omega })t(\omega +m\mathrm{\Omega })J_n(\omega \tau _g)J_m(\omega \tau _g)}`$ (65) Let us assume that the distance between resonances are exactly equal to $`N\mathrm{\Omega }=\mathrm{\Delta }`$ where $`N`$ is an integer. Then substituting the transmission amplitude $`t(\omega )`$into this expression and making the integration over $`\omega `$ one gets: $`P(t)4\alpha {\displaystyle \frac{S}{r^2}}{\displaystyle \underset{n=\mathrm{}}{\overset{\mathrm{}}{}}}n(\omega _0n\mathrm{\Omega })\mathrm{\Gamma }{\displaystyle \frac{\mathrm{}(\omega _0n\mathrm{\Omega })}{E_{det}}}\left[J_n^2[(\omega _0n\mathrm{\Omega })\tau _g]+J_n[(\omega _0n\mathrm{\Omega })\tau _g]J_{n+N}[(\omega _0n\mathrm{\Omega })\tau _g]\mathrm{cos}N\mathrm{\Omega }t\right]`$ (66) Then for the special choice $`N=2`$ assuming that the argument of the Bessel function is typically large one can put $`J_{n+2}(x)J_n(x)`$ for $`x1`$. Then the photo-detection probability can be written as $`P(t)4\alpha {\displaystyle \frac{S}{r^2}}\left(1\mathrm{cos}2\mathrm{\Omega }t\right){\displaystyle \underset{n=\mathrm{}}{\overset{\mathrm{}}{}}}n(\omega _0n\mathrm{\Omega })\mathrm{\Gamma }{\displaystyle \frac{\mathrm{}(\omega _0n\mathrm{\Omega })}{E_{det}}}J_n^2[(\omega _0n\mathrm{\Omega })\tau _g].`$ (67) One can see that the term which is alternating in time in $`P(t)`$ arises due to interference of the different components of the photon wave packet (26). ### III.4 Analysis of intensity correlations In the previous section, we showed that by appropriate filtering of the light signal, it is possible to extract oscillations associated with the past interaction with the GW. While this previous proposal is promising and attractive, one can take an alternative route for the detection of the GW signal: the measurement of intensity–intensity correlations of the photo-detection signal. Consider the correlator $`I(t_1)I(t_2)`$: because of the time modulation due to the presence of the GW, it no longer depends on the variable $`t_1t_2`$ only. Using the definition the intensity operator, one can see qualitatively that the Fourier components of Eq. (43) will lead to a two particle interference effect. The quantity to be measured on the Earth is thus the intensity correlator of the photo-detection signal which is the time ordered product of electrical field operators: $$:\widehat{I}(t_1)\widehat{I}(t_2):dt_1dt_2=\alpha ^2\frac{S^2c^2dt_1dt_2}{E_{det}^2}\widehat{𝐄}^{()}(t_1)\widehat{𝐄}^{()}(t_2)\widehat{𝐄}^{(+)}(t_2)\widehat{𝐄}^{(+)}(t_1)$$ (68) Substituting here the expression for the electrical field operators (36) and performing the quantum average one gets $`:\widehat{I}(t_1)\widehat{I}(t_2):dt_1dt_2=2\alpha ^2{\displaystyle \frac{S^2dt_1dt_2}{r^4}}{\displaystyle }d\omega _1d\omega _2e^{i\omega _2[t_1t_2+\tau _g(\mathrm{sin}\mathrm{\Omega }t_1\mathrm{sin}\mathrm{\Omega }t_2)]}`$ $`\times {\displaystyle \frac{\mathrm{}^2\omega _1\omega _2}{E_{det}^2}}n(\omega _1)n(\omega _2)e^{i\omega _1[t_1t_2+\tau _g(\mathrm{sin}\mathrm{\Omega }t_1\mathrm{sin}\mathrm{\Omega }t_2)]}.`$ (69) This correlator can be rewritten as $$:\widehat{I}(t_1)\widehat{I}(t_2):dt_1dt_2=2\alpha ^2\frac{S^2dt_1dt_2}{r^4}\left|f[t_1t_2+\tau _g(\mathrm{sin}\mathrm{\Omega }t_1\mathrm{sin}\mathrm{\Omega }t_2)]\right|^2,$$ (70) where $$f(t)=\frac{\mathrm{}\omega }{E_{det}}n(\omega )e^{i\omega t}𝑑\omega .$$ (71) The function $`f(t)`$ typically decays exponentially on a time scale given by the inverse of the line width $`\mathrm{\Gamma }_0\mathrm{\Omega }`$. Thus one can safely linearize the dependence of the intensity-intensity correlator with respect to the relative time $`\tau =t_1t_2`$: $`:\widehat{I}(t+\tau )\widehat{I}(t):=2\alpha ^2{\displaystyle \frac{S^2}{r^4}}\left|f[\tau (1+\mathrm{\Omega }\tau _g\mathrm{cos}\mathrm{\Omega }t)]\right|^2`$ (72) where $`\tau =t_1t_2`$ is the relative time and $`t`$ is the total time. Let us now calculate the spectral power, $`S(t)`$ of the photons’ noise defined as $$S(t)=:\widehat{I}(t+\tau )\widehat{I}(t):d\tau $$ (73) Substituting here the expression for photon intensity correlator one can show that $$S(t)=\frac{4\pi \alpha ^2}{1+\mathrm{\Omega }\tau _g\mathrm{cos}\mathrm{\Omega }t}\frac{S^2}{r^4}\frac{(\mathrm{}\omega )^2}{E_{det}^2}n^2(\omega )𝑑\omega $$ (74) One can see from this equation that the intensity correlation contain periodic oscillations – although with a small amplitude – in the total time. Note that contrary to the Fabry-Perot diagnosis, here a rather accurate measurement is implied, as the contribution to other noise sources (due to scattering…) needs to be minimized compared to this periodic signal. In particular, it requires a large time acquisition window to filter out spurious fluctuations. ### III.5 Conclusion Let us now summarize all results obtained through the article and give some estimations and limitations for the proposed gravitational waves detection method. The primary effect we have dealt with is the modulation of the photon phase due to interaction with gravitational field have been found in the geometrical optics approximation and described by eq. (26) for the photon wave function: $$\varphi _k(\xi )=𝒞(\xi )e^{i\omega _k\xi i\omega _k\tau _g\mathrm{sin}\mathrm{\Omega }\xi },$$ (75) where $`\xi =tr/c`$ and $`𝒞(\xi )=\left[1+\mathrm{\Omega }\tau _g\mathrm{cos}\mathrm{\Omega }\xi \right]^1`$ is the pre-exponential factor. This effect of phase modulation leads immediately to well known Shapiro effect of the time delay of the signal propagating near the source of gravitational field. For the case of the source of gravitational waves this time delay is an alternating function of time $`\tau _{ar}=r/c\tau _g\mathrm{sin}\mathrm{\Omega }\tau `$, where $`\tau _{ar}`$ is an arrival time of the signal, $`r/c`$ the traveling time of the signal in flat Minkovski space and $$\tau _g=\frac{h_{xx}(\mathrm{\Lambda })}{\pi }\left(\frac{\mathrm{\Lambda }}{D}\right)^2\mathrm{\Omega }^1.$$ (76) Here $`h_{xx}(\mathrm{\Lambda })`$ is the dimensionless strength of gravitational wave near the GW source, $`D`$ impact parameter, $`\mathrm{\Omega }=2\pi c/\mathrm{\Lambda }`$ is the frequency of gravitational waves. The Shapiro effect in fact seems to be a very promising candidate for gravity wave detection experiment since the time delay (or the additional alternating in time phase) mainly accumulated near the source of gravitational waves where the strength of GW many order of magnitude greater when they reach the Earth. However in the proposed experiment sazhin\_1 it requires a pulsar on the line of sight which passes near the source of gravitational radiation. In this paper we propose a gravitational waves detection method based on the Fabry-Perot interference analysis (or equivalently time correlation measurement) of the light signal from arbitrarily light source (not necessarily a pulsar) passing near the strong source of gravitational radiation. Then the photodetector signal contains an alternating in time component (with GW frequency) proportional to the strength of gravitational radiation near the the GW source (see eq. 56, 61): $$P(t)=P_0(1\mathrm{\Omega }\tau _g\mathrm{cos}\mathrm{\Omega }t),$$ (77) with $`P_0`$ a photodetector signal corresponding to a light coming to a FP filter from distant star. To estimate an effectiveness of the proposed method let us first estimate a brightness of the light source needed to resolve an alternating component in the photo-detection signal. Collecting the signal over long time $`\tau _{\mathrm{obs}}`$ allow to better resolve the alternating component. Assuming that during the observation time $`N_{\mathrm{obs}}`$ photons coming from a distant star hit the photodetector, in the limit of Poissonian statistics, $`\delta N_{\mathrm{obs}}^2=N_{\mathrm{obs}}`$ one has the following limiting requirement: $$\mathrm{\Omega }\tau _g(D)\sqrt{N_{\mathrm{obs}}}1.$$ (78) in the extreme case where the light passes in immediate proximity ($`D\mathrm{\Lambda }`$) to the GW source one has the condition: $$h(\mathrm{\Lambda })\sqrt{N_{\mathrm{obs}}}1.$$ (79) Then assuming that the distant star emits in the frequency band $`\mathrm{\Delta }\omega `$ and that the diameter of the telescope equals to $`d`$ one has: $$N_{\mathrm{obs}}L\frac{d^2}{R^2}\frac{\mathrm{\Gamma }}{\mathrm{\Delta }\omega }\frac{\tau _{\mathrm{obs}}}{\mathrm{}\omega _0},$$ (80) where $`L`$ is the brightness of the distant star, $`R`$ is the distance to the Earth. The most promising candidates of gravitational waves sources appropriate for proposed detection method are the periodic sources of gravitational radiation such as close binaries at the final stage of evolution or a asymmetric rotating neutron stars, allowing to collect the signal long enough time to resolve the tiny gravitational wave modulation on the stochastic Poisson noise. Let us give some estimations, assuming the total mass, $`m`$ of binary system of the order of the mass of the Sun, and rotating with a frequency $`\nu 10^2`$ Hz. Then the dimensionless strength of gravitational waves at the distance $`r`$ from the emitter, according to quadrupole approximation, is given by $$h4\frac{G^{5/3}}{c^4}\frac{1}{r}m^{5/3}\nu ^{2/3}$$ (81) assuming that the role of the distant star plays one of the components of the binary system $`r=(Gm/\nu ^2)^{1/3}`$ one has $`h(r)4(Gm\nu )^{4/3}/c^410^9`$. In order to resolve this modulation one needs about $`N_{\mathrm{obs}}10^{18}`$ photons coming to a photodetector during $`\tau _{\mathrm{obs}}`$. Assuming the the typical distance from the binary system to the Earth of the order $`1`$ kpa with $`d10m`$, $`\omega _010^{14}`$ Hz and $`\mathrm{\Gamma }/\mathrm{\Delta }\omega 1`$, $`\tau _{\mathrm{obs}}10^3\nu ^110^5s`$ one has the required brightness of the binary component, according to eq. (80)of the order $`L10^5L_\mathrm{S}`$ where $`L_\mathrm{S}`$ is the brightness of the Sun. Of course the requirement to have such a bright component of the binary system makes the observation of the proposed effect rather problematic in practice. In reality the situation with binary systems is even worse due to the presence of the Doppler effect for the light emitted from one of the rotating components. In fact the Doppler effect results in the same frequency modulation of the emitted light as the gravitational waves modulation and the amplitude of the frequency modulation due to the Doppler effect is of the order of $$\left(\frac{\mathrm{\Delta }\omega }{\omega }\right)_{\mathrm{Doppler}}\left(\frac{v}{c}\right)^2\frac{(Gm\nu )^{1/3}}{c}\frac{\nu }{c}.$$ (82) For parameters described above it results in an effect of the order $`10^5`$ that is five orders of magnitude greater than the frequency modulation due to gravitational waves. The situation however could be much better for very close neutron star binaries at the last stage of evolution. In this case the rotational frequency can be of the order of $`10^2`$ Hz and results in a dimensionless strength of emitted gravitational waves of the order $`h(\mathrm{\Lambda })10^6`$. Assuming that the distance to the binary neutron star system is of the order $`1`$ kpa and that there is a bright distant star situated on the line of sight to the binary from the Earth with impact parameter $`D\mathrm{\Lambda }`$ and collecting the signal during the time $`\tau _{\mathrm{obs}}10^3\nu ^110s`$ one has the following lower limit for the brightness of distant star $`L10^2L_\mathrm{S}`$. ### III.6 Acknowledgments. We acknowledge discussions to M. Sazhin, A. Starobinsky and K. Bayandin. A. L. thanks the Ecole-Normale Landau Institute agreements and CNRS for his stay at the Centre de Physique Théorique. He also acknowledges the Forschungszentrum Juelich for financial support within the framework of the Landau Program. We acknowledge financial support from the Russian Science Support Foundation, the Russian Ministry of Science, and the program ‘Quantum Macrophysics’ of the RAS.
warning/0506/hep-ph0506183.html
ar5iv
text
# 1 Mirror Universe Models ## 1 Mirror Universe Models Mirror universe models, with hidden sector matter and forces identical to that of the visible world, but interacting with the latter only via gravity, have been proposed in the context of neutrino physics (specifically by attempts to understand the nature of a sterile neutrino) and in the context of superstring/M-theory . As a means to reconcile the constraints of big bang nucleosynthesis, the reheating temperature of the mirror universe after inflation was postulated to be lower than that in the observable universe . From this it was shown that the asymmetric reheating can be related to a difference of the electroweak symmetry breaking scales in the two sectors, as required for a mirror solution to the neutrino puzzle. In such models it was shown that the baryon asymmetry is greater in the mirror universe than in the observable universe and that the mirror baryons could provide the dominant dark matter in the bulk universe . While heterotic strings (M-theory) offer a mirror universe scenario in ten (eleven) dimensions, with the well-known $`E_8`$ gauge symmetry for both observable and hidden sectors (9-branes), string-derived quasi-realistic three generation mirror universe models in four-dimensions have, to our knowledge, not been constructed. Rather, in typical quasi-realistic string models, the ten-dimensional mirror symmetry is broken through compactification to four dimensions. In bosonic lattice/orbifold construction differentiation of observable and hidden gauge groups (and, thus, of matter representations) is typically a result of asymmetric Wilson loop effects and spin embedding. In corresponding free fermionic models, mirror symmetry breaking generally results from asymmetric boundary conditions between observable and hidden sector worldsheet fermions. An investigation into string/M-derived mirror models is underway and will be reported in an upcoming paper . Contrastingly, in this letter we wish to discuss an obstacle to mirror model construction that can develop in weakly coupled heterotic strings. This obstacle, nonetheless, produces an interesting physical effect in its own right. In this letter we will demonstrate how, under a specific set of conditions, mirror symmetry breaking will necessarily occur in models with symmetric boundary conditions. We will investigate the implications of this for initially mirror symmetric models. During our string/M-derived mirror universe investigation, we found that even when symmetric worldsheet fermion boundary conditions are imposed, breaking of mirror symmetry is sometimes mandated by an unavoidable asymmetry imposed by the GSO projections. In this note we show, through two example models, how GSO projections can necessitate mirror symmetry breaking of observable ($`O`$) and hidden sector ($`H`$) Pati-Salam gauge groups, $`[SU(4)_CSU(2)_LSU(2)_R]^O[SU(4)_CSU(2)_LSU(2)_R]^H.`$ (1.1) In our first example, the GSO projections reduce the observable sector gauge group to $`[SU(4)_CSU(2)_L]^O,`$ (1.2) by transferring $`SU(2)_R^O`$ to the hidden sector. In the process the $`[SU(4)_CSU(2)_L]^H`$ (1.3) subgroup of the hidden sector Pati-Salam gauge symmetry is enhanced to $`[SO(10)]^H.`$ (1.4) Additionally, the initial shadow $`(S)`$ sector (corresponding to charges possibly carried by both observable and hidden states) gauge group, $`[SU(2)^3SU(3)U(1)^7]^S`$, absorbs $`[SU(2)_R]^H`$ to become $`[SU(2)^3SU(5)U(1)^6]^S`$ (see Table B.2). Our second example differs from the first in some of the elements of its GSO projections matrix. For model 2, the GSO projections reduce the observable sector Pati-Salam gauge group to $`[SU(4)_CSU(2)_R]^O`$ by alternately transferring $`SU(2)_L^O`$ to the hidden sector. The hidden sector $`[SU(4)_CSU(2)_R]^H`$ subgroup of the Pati-Salam group is similarly enhanced to $`[SO(10)]^H`$. In this version, the shadow sector gauge group remains of rank 12 and does not absorb $`[SU(2)_L]^H`$ (see Table C.2). In Section 2 we briefly review free fermionic construction and the related form of the GSO projections. In Section 3 we then review the NAHE-based $`\text{ZZ}_2\times \text{ZZ}_2`$ models , the class to which our example models belong (see Table A.1). Following this, in Section 4 we introduce the mirror-symmetric free fermion worldsheet boundary conditions for our two models (see Tables A.2 and A.3). We then demonstrate how an unavoidable asymmetry that arises between certain GSO projection matrix elements in our models produces mirror symmetry breaking between observable and hidden sector states. We study viable hypercharge definitions in Section 5. Our investigation of the unavoidable observable/hidden sector mirror-breaking effect in our models is then summarized in Section 6. We present the gauge group and matter states for our first model in Tables B.2 and B.3, respectively, and those for our second model in Tables C.2 and C.3, respectively. ## 2 Free Fermionic Models A free fermionic heterotic string model is specified by two objects . The first is a $`p`$-dimensional basis set of free fermionic boundary vectors $`\{𝐕_i`$, $`i=1`$, … , $`p\}`$. Each vector $`𝐕_i`$ has 64 components, $`1<V_i^m1`$, $`m=1`$, … , 64, with the first 20 components specifying boundary conditions for the 20 real free fermions representing worldsheet degrees of freedom for the left-moving supersymmetric string, and the later 44 components specifying boundary conditions for the 44 real free fermions representing worldsheet degrees of freedom for the right-moving bosonic string. (Components of $`𝐕_i`$ for complex fermions are double-counted.) Modular invariance dictates that the basis vectors, $`𝐕_i`$, span a finite additive group $`\mathrm{\Xi }=\{_{n_i=1}^{N_i1}_{i=1}^pn_i𝐕_i\}`$, with $`N_i`$ the lowest positive integer such that $`N_i𝐕_i=\mathrm{𝟎}(\mathrm{mod}\mathrm{\hspace{0.17em}2})`$. $`𝐕_1=\mathrm{𝟏}`$ (a 64-component unit vector) must be present in all models. In a given sector $`𝜶`$$`a_i𝐕_i(\mathrm{mod}\mathrm{\hspace{0.17em}2})\mathrm{\Xi }`$, with $`a_i\{0,\mathrm{\hspace{0.17em}1},\mathrm{},N_i1\}`$, a worldsheet fermion $`f_m`$ transforms as $`f_m\mathrm{exp}\{\pi \alpha _m\}f_m`$ around non-contractible loops on the worldsheet. Boundary vector components for real fermions are thus limited to be either 0 or 1, whereas boundary vector components for complex fermions can be rational . The second object necessary to define a free fermionic model (up to vacuum expectation values (VEVs) of fields in the effective field theory) is a $`p\times p`$-dimensional matrix $`𝐤`$ of rational numbers $`1<k_{i,j}1`$, $`i,j=1`$, …, $`p`$, that determine the GSO operators for physical states. The $`k_{i,j}`$ are related to the phase weights $`C\left(\genfrac{}{}{0pt}{}{𝐕_i}{𝐕_j}\right)`$ appearing in a model’s one-loop partition function $`Z`$: $`C\left({\displaystyle \genfrac{}{}{0pt}{}{𝐕_i}{𝐕_j}}\right)=(1)^{s_i+s_j}\mathrm{exp}(\pi ik_{j,i}\frac{1}{2}𝐕_i𝐕_j),`$ (2.1) where $`s_i`$ is the 4-dimensional spacetime component of $`𝐕_i`$. (The inner product of boundary (or charge) vectors is lorentizian, taken as left-movers minus right movers. Contributions to inner products from real fermion boundary components are weighted by a factor of $`\frac{1}{2}`$ compared to contributions from complex fermion boundary components.) The phase weights $`C\left(\genfrac{}{}{0pt}{}{𝜶}{𝜷}\right)`$ for general sectors $`𝜶={\displaystyle \underset{j=1}{\overset{p}{}}}a_j𝐕_j\mathrm{\Xi },𝜷={\displaystyle \underset{i=1}{\overset{p}{}}}b_i𝐕_i\mathrm{\Xi }`$ (2.2) can be expressed in terms of the components in the $`p\times p`$-dimensional matrix $`𝐤`$ for the basis vectors: $`C\left({\displaystyle \genfrac{}{}{0pt}{}{𝜶}{𝜷}}\right)=(1)^{s_\alpha +s_\beta }\mathrm{exp}\{\pi i{\displaystyle \underset{i,j}{}}b_i(k_{i,j}\frac{1}{2}𝐕_i𝐕_j)a_j\}.`$ (2.3) Modular invariance simultaneously imposes constraints on the basis vectors $`𝐕_i`$ and on components of the GSO projection matrix $`𝐤`$: $`k_{i,j}+k_{j,i}`$ $`=`$ $`\frac{1}{2}𝐕_i𝐕_j(\mathrm{mod}\mathrm{\hspace{0.17em}2})`$ (2.4) $`N_jk_{i,j}`$ $`=`$ $`0(\mathrm{mod}\mathrm{\hspace{0.17em}2})`$ (2.5) $`k_{i,i}+k_{i,1}`$ $`=`$ $`s_i+\frac{1}{4}𝐕_i𝐕_i(\mathrm{mod}\mathrm{\hspace{0.17em}2}).`$ (2.6) The dependence upon the $`k_{i,j}`$ can be removed from equations (2.4-2.6), after appropriate integer multiplication, to yield three constraints on the $`𝐕_i`$: $`N_{i,j}𝐕_i𝐕_j=0(\mathrm{mod}\mathrm{\hspace{0.17em}4})`$ (2.7) $`N_i𝐕_i𝐕_i=0(\mathrm{mod}\mathrm{\hspace{0.17em}8})`$ (2.8) $`\mathrm{The}\mathrm{number}\mathrm{of}\mathrm{real}\mathrm{fermions}\mathrm{simultaneously}\mathrm{periodic}`$ $`\mathrm{for}\mathrm{any}\mathrm{three}\mathrm{basis}\mathrm{vectors}\mathrm{is}\mathrm{even}.`$ (2.9) $`N_{i,j}`$ is the lowest common multiple of $`N_i`$ and $`N_j`$. (2.9) applies when two or more of three the basis vectors are identical. Thus, each basis vector must have an even number of real periodic fermions. The physical massless states in the Hilbert space of a given sector $`𝜶`$$`\mathrm{\Xi }`$, are obtained by acting on the vacuum with bosonic and fermionic operators and by applying generalized GSO projections. The $`U(1)`$ charges for the Cartan generators of the unbroken gauge group are in one to one correspondence with the $`U(1)`$ currents $`f_m^{}f_m`$ for each complex fermion $`f_m`$, and are given by: $$Q_m^𝜶=\frac{1}{2}\alpha _m+F_m^𝜶,$$ (2.10) where $`\alpha _m`$ is the boundary condition of the worldsheet fermion $`f_m`$ in the sector $`𝜶`$, and $`F_m^𝜶`$ is a fermion number operator counting each mode of $`f_m`$ once and of $`f_m^{}`$ minus once. Pseudo-charges for non-chiral (i.e., with both left- and right-moving components) real Ising fermions $`f_m`$ can be similarly defined, with $`F_m`$ counting each real mode $`f`$ once. For periodic fermions, $`\alpha _m=1`$, the vacuum is a spinor representation of the Clifford algebra of the corresponding zero modes. For each periodic complex fermion $`f_m`$ there are two degenerate vacua $`|+,|`$ , annihilated by the zero modes $`(f_m)_0`$ and $`(f_m^{})_0`$ and with fermion numbers $`F_m^𝜶=0,1`$, respectively. The boundary basis vectors $`𝐕_j`$ generate the set of GSO projection operators for physical states from all sectors $`𝜶`$$`\mathrm{\Xi }`$. In a given sector $`𝜶`$, the surviving states are those that satisfy the GSO equations imposed by all $`𝐕_j`$ and determined by the $`k_{j,i}`$’s: $`𝐕_j𝐅^𝜶=\left({\displaystyle \underset{i}{}}k_{j,i}a_i\right)+s_j\frac{1}{2}𝐕_j𝜶(\mathrm{mod}\mathrm{\hspace{0.17em}2}),`$ (2.11) or, equivalently, $`𝐕_j𝐐^𝜶=\left({\displaystyle \underset{i}{}}k_{j,i}a_i\right)+s_j(\mathrm{mod}\mathrm{\hspace{0.17em}2}).`$ (2.12) For a given set of basis vectors, the independent GSO matrix components are $`k_{1,1}`$ and $`k_{i,j}`$, for $`i>j`$. This GSO projection constraint, when combined with equations (2.4-2.6) form the free fermionic re-expression of the even, self-dual modular invariance constraints for bosonic lattice models. ## 3 $`\text{ZZ}_2\times \text{ZZ}_2`$ NAHE-Based Models The $`\text{ZZ}_2\times \text{ZZ}_2`$ NAHE-based models have contributed much to understanding regions of the parameter space of weakly-coupled (quasi)-realistic string models, including GUT’s , semi-GUT’s , near-MSSM’s , and MSSM’s . They have advanced knowledge of general features of classes of string model, both perturbative and non-perturbative properties that likely still persist after M-theory embedding. The NAHE set consists of the five basis vectors, $`\mathrm{𝟏}`$, $`𝐒`$, and $`𝐛_i`$, $`i=1`$ to 3, given in Table A.1. The NAHE set breaks the observable sector $`E_8`$ gauge symmetry to $`SO(10)U(1)^3`$ but does not affect the hidden sector $`E_8`$. The six compactified directions, represented by six pairs of left- and right-moving real fermions $`\{y,\omega |\overline{y},\overline{\omega }\}^{1,\mathrm{},6}`$, generate an additional $`SO(12)`$ shadow sector (S) gauge symmetry. The NAHE set breaks this to $`SO(4)^3`$. However, each $`U(1)_i`$ combines with a corresponding $`SO(4)_i`$ to form an enhanced $`SO(6)_i`$ symmetry. Thus, the gauge group of the NAHE set is $`[SO(10)SO(6)^3]^OE_8^H`$ with $`N=1`$ spacetime supersymmetry. The observable sector matter is 48 spinorial $`\mathrm{𝟏}6`$’s of $`SO(10)`$, with sixteen of these from each sector $`𝐛_1`$, $`𝐛_2`$ and $`𝐛_3`$. The $`U(1)_i`$, for $`i=1`$ to 3, act as generation charges for the three sectors $`𝐛_1`$, $`𝐛_2`$ and $`𝐛_3`$, respectively. The sixteen copies of $`\mathrm{𝟏}6`$’s of $`SO(10)`$ from a given $`𝐛_i`$ are formed from 4 copies of $`\mathrm{𝟒}`$’s of $`SO(6)_i`$. To obtain three generation models, the $`SO(6)^3`$ must be broken to no more than the initial $`U(1)^3`$. Reduction of the $`SO(10)`$ and of the $`SO(6)^3`$ can result from additional boundary basis vectors that will break $`SO(10)`$ to one of its subgroups, $`SU(5)\times U(1)`$, $`SO(6)\times SO(4)`$ or $`SU(3)\times SU(2)\times U(1)^2`$. The additional sectors must simultaneously reduce the number of generations to three, one from each of the sectors $`𝐛_1`$, $`𝐛_2`$ and $`𝐛_3`$. The detailed phenomenological properties of the various NAHE-based three generation models can differ significantly. In particular, the properties strongly depend on the boundary components of the non-NAHE basis vectors for the fermions representing the six compactified directions . ## 4 Symmetry Breaking of Mirror Models Mirror models with matching observable and hidden sector symmetries and states may be created from NAHE-based models by adding mirror basis vectors, $`𝐛_1^{^{}}`$, $`𝐛_2^{^{}}`$ and $`𝐛_3^{^{}}`$, as defined in Table A.2. Due to the symmetry between $`𝐛_i`$ and $`𝐛_i^{^{}}`$, these mirror vectors break the hidden sector $`E_8`$ in the same manner as the NAHE set breaks the observable sector $`E_8`$ into $`SO(10)U(1)^3`$. Each $`𝐛_i^{^{}}`$ produces 16 copies of 16’s of the hidden sector $`SO(10)`$. The right-moving components of the $`𝐛_i`$ and the $`𝐛_i^{^{}}`$ basis vectors that are simultaneously non-zero form a subset of $`\{y,\omega |\overline{y},\overline{\omega }\}^{1,\mathrm{},6}`$. That is, $`𝐛_i`$ and $`𝐛_i^{^{}}`$ states will both carry some $`\{y,\omega |\overline{y},\overline{\omega }\}^{1,\mathrm{},6}`$ “shadow sector” charges. Within the NAHE set the only two bosonic sectors that can produce gauge states are $`\mathrm{𝟎}`$ and $`\mathrm{𝟏}+𝐛_1+𝐛_2+𝐛_3`$. All observable and “shadow sector” $`SO(n)`$ and $`U(1)`$ generators, along with the hidden sector $`\mathrm{𝟏}20`$ rep of $`SO(16)E_8`$, originate in $`\mathrm{𝟎}`$, while the hidden sector $`\mathrm{𝟏}28`$ rep of $`SO(16)E_8`$ originates in $`\mathrm{𝟏}+𝐛_1+𝐛_2+𝐛_3`$. The GSO projections from the mirror sectors $`𝐛_1^{}`$, $`𝐛_2^{}`$, and $`𝐛_3^{}`$ remove the $`\mathrm{𝟏}28`$ rep of $`SO(16)`$, reducing the hidden sector $`E_8`$ symmetry to $`SO(16)`$. Further, the mirror sectors reduce the hidden sector $`SO(16)`$ to an $`SO(10)U(1)^3`$ symmetry, matching the observable sector. (Similarly the GSO projections of $`𝐛_1`$, $`𝐛_2`$, and $`𝐛_3`$ remove any contribution to the observable gauge group from the $`\mathrm{𝟏}+𝐛_1^{}+𝐛_2^{}+𝐛_3^{}`$ sector.) The hidden sector $`U(1)^3`$ generation charges and the observable sector $`U(1)^3`$ generation charges combine with the shadow sector charges in a like manner. Simultaneously, the $`𝐛_i^{}`$ sector GSO’s reduce the $`SO(4)^3`$ shadow sector symmetry to $`U(1)^6`$. The net result is a $`SU(3)SU(2)^3U(1)^7`$ shadow sector symmetry, whose charges are carried by matter representations of the observable $`SO(10)`$ and of the hidden $`SO(10)`$. The additional $`SU(3)SU(2)^3`$ generators originate in several additional massless gauge sectors formed from linear combinations of $`\mathrm{𝟏}`$, $`𝐛_i`$, and $`𝐛_i^{}`$. All of these sectors have massless vacua that only carry $`\overline{\eta }_i`$, $`\overline{\eta }_i^{}`$, and $`\overline{y}`$, $`\overline{\omega }`$ shadow charges. While most massless gauge states from these sectors are projected out, a few are not. Those that survive combine observable sector $`\overline{\eta }`$-charges, hidden sector $`\overline{\eta }^{}`$-charges, and shadow sector $`\overline{y},\overline{\omega }`$-charges (see Tables B.2 and C.2). Thus, at this stage mirror symmetry still exists, as should be expected. Note in particular that $`𝐤`$ is invariant under exchange of $`𝐛_i`$ with co responding $`𝐛_i^{}`$. To break each of the $`SO(10)`$ to their corresponding Pati-Salem $`SU(4)_CSU(2)_LSU(2)_R`$, two additional sectors $`𝐚`$ and $`𝐚^{}`$ are added to the model, where $`𝐚`$ and $`𝐚^{}`$ are mirror sectors, as shown in in Table A.2. The set of 10 basis vectors $`\{\mathrm{𝟏},𝐒,𝐛_1,𝐛_2,𝐛_3,𝐛_1^{},𝐛_2^{},𝐛_3^{},𝐚,𝐚^{}\}`$, produce a model with mirror symmetry boundary conditions for the observable and hidden sectors. However, with the addition of $`𝐚`$ and $`𝐚^{}`$, an unavoidable asymmetry develops among the GSO projection components: Although all new degrees of freedom of $`k_{𝜶,𝜷}`$, for $`𝜶`$$`\{𝐚,𝐚^{}\}`$ and $`𝜷`$$`\{\mathrm{𝟏},𝐒,𝐛_1,𝐛_2,𝐛_3,𝐛_1^{},𝐛_2^{},𝐛_3^{}\}`$, are chosen to be invariant under simultaneous exchange of primed and un-primed sectors for both $`𝜶`$ and $`𝜷`$, symmetry breaking between $`k_{𝐚,𝐚^{}}`$ and $`k_{𝐚^{},𝐚}`$ occurs automatically. Only one of $`k_{𝐚,𝐚^{}}`$ and $`k_{𝐚^{},𝐚}`$ is a degree of freedom; the other is specified by (2.4). As Table A.3 shows, $`𝐚𝐚^{}=10`$, which from (2.4) yields $`k_{𝐚,𝐚^{}}+k_{𝐚^{},𝐚}`$ $`=1(\mathrm{mod}\mathrm{\hspace{0.17em}2}).`$ (4.1) Thus, $`k_{𝐚,𝐚^{}}`$ and $`k_{𝐚^{},𝐚}`$ cannot be equal (mod 2); either $`k_{𝐚,𝐚^{}}=1`$ and $`k_{𝐚^{},𝐚}=0`$ or vice versa (since all components of $`𝐚`$ and $`𝐚^{}`$ are either anti-periodic or periodic). The addition of $`𝐚`$ and $`𝐚^{}`$ to the model generates several new massless gauge sectors of the form $`𝐚+𝐚^{}+\mathrm{}`$. However, in both model variations presented herein, the GSO projections remove all possible gauge states from such sectors, except those coming from $`𝐚+𝐚^{}`$. The $`𝐚+𝐚^{}`$ sector has all anti-periodic components except for four periodic associated with the two complex fermions generating the observable $`SO(4)=SU(2)_LSU(2)_R`$ and the two complex fermions generating the hidden sector $`SO(4)=SU(2)_LSU(2)_R`$. Thus, in the $`𝐚+𝐚^{}`$ sector, massless gauge states require one anti-periodic fermionic (with $`Q=\pm 1`$) excitation. The GSO projection from $`𝐚`$ acts on observable $`SO(4)`$ spinors while $`𝐚^{}`$ acts on hidden $`SO(4)`$ spinors. Since $`k_{𝐚,𝐚^{}}`$ and $`k_{𝐚,𝐚^{}}`$ differ by 1 (mod 2), so do $`k_{𝐚,𝐚+𝐚^{}}`$ and $`k_{𝐚^{},𝐚+𝐚^{}}`$. Thus, a state in the $`𝐚+𝐚^{}`$ sector survives both the $`𝐚`$ sector GSO projections and $`𝐚^{}`$ sector GSO projections if and only if its observable and hidden $`SO(4)`$ spinors have opposite chirality. That is, the net number of down spins among the four spinors must be odd, implying that an $`𝐚+𝐚^{}`$ gauge state will either carry observable $`SU(2)_LSO(4)`$ charge and hidden $`SU(2)_RSO(4)`$ charge or vice-versa. For model 1, the additional $`k_{𝐛_i^{(^{})},𝐚}`$ and $`k_{𝐛_i^{(^{})},𝐚^{}}`$ require $`𝐚+𝐚^{}`$ states to always have an even number of observable $`SO(4)`$ $`\frac{1}{2}`$ spins and an odd number of hidden $`SO(4)`$ $`\frac{1}{2}`$ spins, linking observable $`SU(2)_R`$ with hidden $`SU(2)_L`$. The remaining GSO projections on $`𝐚+𝐚^{}`$ gauge states require the $`Q=\pm 1`$ anti-periodic fermion excitation charge to be from the hidden sector $`SO(6)SU(4)`$. Thus, the surviving $`𝐚+𝐚^{}`$ simple root connects observable $`SU(2)_R`$ roots with the hidden sector $`SO(6)`$ and $`SU(2)_L`$ roots, thereby regenerating a hidden sector $`SO(10)`$ (see Table B.2): $`[SU(2)_R]^O[SO(6)SU(2)_L]^H[SO(10)]^H.`$ (4.2) In addition, a sector $`1+𝐛_1+𝐛_1^{}+𝐚`$ gauge state with eight complex spinors links the shadow gauge states with the hidden sector $`SU(2)_R`$, increasing the shadow sector gauge symmetry, $`[SU(3)SU(2)^3U(1)^7]^S(SU(2)_R)^H[SU(5)SU(2)^3U(1)^6]^S.`$ (4.3) The final model 1 gauge group is, therefore, $`[SU(4)SU(2)_L]^O[SU(5)SU(2)^3U(1)^6]^S[SO(10)]^H.`$ (4.4) While maintaining symmetry under exchange of primed and un-primed components of $`𝐤`$, model 2 differs from model 1 in some choices of $`k_{𝐚,\beta }`$ and $`k_{𝐚^{},\beta }`$, for $`𝜷`$$`\{\mathrm{𝟏},𝐒,𝐛_1,𝐛_2,𝐛_3,𝐛_1^{},𝐛_2^{},𝐛_3^{}\}`$ (see Table C.1). Model 2 GSO projections require an odd number of observable $`SO(4)`$ $`\frac{1}{2}`$ spins and an even number of hidden $`SO(4)`$ $`\frac{1}{2}`$ spins in the $`𝐚+𝐚^{}`$ sector. This link observable $`SU(2)_L`$ reps with hidden $`SU(2)_R`$ reps. The remaining GSO projections again require the $`Q=\pm 1`$ anti-periodic fermion excitation charge of an $`𝐚+𝐚^{}`$ simple root to be from the hidden sector $`SO(6)`$. Thus, for model 2 the gauge boson from $`𝐚+𝐚^{}`$ again regenerates a hidden sector $`SO(10)`$: $`[SU(2)_L]^O[SO(6)SU(2)_R]^H[SO(10)]^H.`$ (4.5) However, in model 2, no additional gauge state is produced to mix the shadow sector with the hidden sector $`SU(2)_L`$. The final model 2 gauge group is, therefore, $`[SU(4)SU(2)_R]^O[SU(3)\times SU(2)^3\times U(1)^7]^S[SO(10)SU(2)_L]^H.`$ (4.6) (An exchange of definitions of left and right-handedness in the observable sector and of related SM states transforms this into $`[SU(4)SU(2)_L]^O[SU(3)\times SU(2)^3\times U(1)^7]^S[SO(10)SU(2)_L]^H.)`$ (4.7) ## 5 Hypercharge Definitions An important issue for these models is whether an acceptable definition of hypercharge can be found, since the conventional hypercharge is missing either its equivalent $`T_3^R`$ contribution in model 1 or its equivalent $`T_3^L`$ contribution in model 2. In standard NAHE-based models the hypercharge is formed as $`Y=\frac{1}{3}\stackrel{~}{Q}_C+\frac{1}{2}\stackrel{~}{Q}_L,`$ (5.1) (for $`Y(Q_L)=\frac{1}{3}`$ normalization), where $`\stackrel{~}{Q}_C=_{m=1}^3Q_{\overline{\psi }^m}`$ is the associated charge trace of $`U(1)_C[\overline{\psi }^1\overline{\psi }^1+\overline{\psi }^2\overline{\psi }^2+\overline{\psi }^3\overline{\psi }^3]`$ and $`\stackrel{~}{Q}_L=_{m=4}^5Q_{\overline{\psi }^m}`$ is the associated charge trace of $`U(1)_L[\overline{\psi }^4\overline{\psi }^4+\overline{\psi }^5\overline{\psi }^5]`$. Since $`SU(4)_CSU(3)_CU(1)_{BL}`$, $`\frac{1}{3}\stackrel{~}{Q}_C=\stackrel{~}{Q}_{BL}`$, as Table 4.a indicates. Similarly, since $`SO(4)=SU(2)_LSU(2)_RSU(2)_LU(1)_L`$, $`\frac{1}{2}\stackrel{~}{Q}_L=2T_3^R`$. Thus, $`Y=\stackrel{~}{Q}_{BL}+2T_3^R,`$ (5.2) which yields electromagnetic charge $`\stackrel{~}{Q}_{EM}=T_3^L+\frac{1}{2}Y=T_3^L+T_3^R+\frac{1}{2}\stackrel{~}{Q}_{BL}.`$ (5.3) Hence, model 1 requires a replacement for $`T_3^R`$, while model 2 needs a replacement for $`T_3^L`$. Under $`SU(4)_CSU(2)_L`$ the SM $`SU(3)_CSU(2)_L`$ left handed reps combine into * $`Q_L=(\mathrm{𝟑},\mathrm{𝟐})_{2T_3^R=0}L_L=(\mathrm{𝟏},\mathrm{𝟐})_{2T_3^R=0}(QL)_L=(\mathrm{𝟒},\mathrm{𝟐})_{2T_3^R=0}`$, * $`e_L^c=(\mathrm{𝟏},\mathrm{𝟏})_{2T_3^R=1}d_L^c=(\mathrm{𝟑},\mathrm{𝟏})_{2T_3^R=1}(d^ce^c)_L=(\mathrm{𝟒},\mathrm{𝟏})_{2T_3^R=1}`$, * $`\nu _L^c=(\mathrm{𝟏},\mathrm{𝟏})_{2T_3^R=1}u_L^c=(\mathrm{𝟑},\mathrm{𝟏})_{2T_3^R=1}(u^c\nu ^c)_L=(\mathrm{𝟒},\mathrm{𝟏})_{2T_3^R=1}`$. Model 1 contains three generations of pairs of $`(q^cl^c)_i^n=(\mathrm{𝟒},\mathrm{𝟏})`$ states (with generation index specified by $`i=1`$ to $`3`$, pair element index specified by $`n=1,2`$, and left-handed index L implicit), where $`(q^cl^c)`$ denotes either $`(d^ce^c)`$ or $`(u^c\nu ^c)`$ (see Table B.3). These states are also respective doublets under the three generation $`SU(2)_i`$ of the shadow sector. As discussed, for a three generation model, the $`SU(2)_i`$ must be broken to the generational $`U(1)_i`$ of standard NAHE models by additional GSO projections from further sectors. When each $`SU(2)_i`$ is broken to $`U(1)_i`$ in this manner, one component of each $`SU(2)_i`$ doublet is also projected out. If these additional GSO projections can be chosen such that the up-spin component of $`SU(2)_i`$ for $`(q^cl^c)_i^{n=1}`$ survives along with the down-spin component of $`SU(2)_i`$ for $`(q^cl^c)_i^{n=2}`$, then for $`Y=\stackrel{~}{Q}_{BL}+2(T_3)_i`$ the $`(q^cl^c)_i^{n=1}`$ become the $`(d^ce^c)_i`$ states and the $`(q^cl^c)_i^{n=2}`$ become the $`(u^c\nu ^c)_i`$ states. Under $`LR`$ exchange, the same process can be applied to create a consistent three generation hypercharge for model 2. For both models, this is the only possible choice for $`(q^cl^c)`$ hypercharge, since in each model the extra Abelian charges carried by the $`(q^cl^c)_i^n`$ are independent of the index $`n`$ for each generation $`i`$. That is, no hypercharge definition involving only the extra $`U(1)_k`$ could yield valid hypercharge for both $`(d^ce^c)`$ and $`(u^c\nu ^c)`$ reps. For model 1, this posses a difficulty for the MSSM higgs. In model 1, the only additional $`SU(2)_L`$ doublets are singlets under all $`SU(2)_i`$. These are the pairs of states $`h_i^n`$ ($`n=1,`$ $`2`$) and the more exotic $`H_i^n`$, which are also 5 reps of $`SU(5)^S`$. For a given generation $`i`$, the extra $`U(1)_k`$ charges are independent of the index $`n`$. Thus, no hypercharge definition could yield both a $`(Y=1)`$-charged up-higgs and a $`(Y=+1)`$-charged down-higgs from an $`h_i^{n=1}`$ and $`h_i^{n=2}`$ pair. Instead, a hypercharge definition is required such that the $`U(1)_k`$ hypercharge contribution to $`QL`$ and $`ql`$ states is zero, while it is +1 for at least one $`h_i^n`$ pair and -1 for at least one other $`h_i^{}^n`$ pair. Applying the six $`QL`$ and $`(q^cl^c)`$ constraints prevents any $`U(1)_k`$ from appearing in a general hypercharge definition. Thus, model 1 cannot provide a suitable definition for hypercharge, unless the Cartan subalgebra of $`SU(5)^S`$ can contribute to the hypercharge of $`H_i^n`$ components after $`SU(5)^S`$ is broken. In contrast, the MSSM higgs of model 2 come in the standard $`h_i`$ and $`\overline{h}_i`$ pairs for each generation i. However, each $`h_i`$ and $`\overline{h}_i`$ is also an $`SU(2)_i`$ doublet (see Table C.3). Thus, if under $`SU(2)_i`$ breaking by GSO projections from additional sectors, the $`SU(2)_i`$ up-spin component of $`h_i`$ survives and the $`SU(2)_i`$ down-spin component of $`\overline{h}_i`$ (or vice-versa) then $`h_i`$ becomes down-higgs and $`\overline{h}_i`$ becomes up-higgs (or vice versa). Thus, a viable hypercharge definition for model 2 is $`Y=\frac{1}{2}\stackrel{~}{Q}_{BL}+{\displaystyle \underset{i=1}{\overset{3}{}}}T_3^i.`$ (5.4) This would produce generational higgs pairs, which is a common occurrence in NAHE-based models. This often provides for mass hierarchy between generations since the physical higgs usually becomes a weighted (by several orders of magnitude) linear combination of the generational higgs. MSSM matter states then couple differently by generation to the physical higgs, producing a large mass hierarchy, even when all mass couplings in the superpotential are third order. Models 1 and 2 both contain an anomalous $`U(1)`$ and it is unlikely that additional basis vectors would remove the anomaly from either model . In fact, additional sectors generally increase the anomaly. For model 2 the charge traces of the seven $`U(1)_k`$ are $`\mathrm{Tr}𝐐=(0,144,96,0,0,192,0).`$ (5.5) (as can be computed from Table C.3). The anomaly may be rotated into a single $`U(1)_A`$, $`U(1)_A=[3Q_2+2Q_34Q_6]`$ (5.6) for which the trace is 1392. The orthogonal $`U_2^{}=2Q_2+3Q_3`$ (5.7) $`U_3^{}=3Q_2+2Q_3+(13/4)Q_6.`$ (5.8) become non-anomalous (traceless). Model 2 is another example for which non-Abelian fields must necessarily take on VEVs to cancel the Fayet-Iliopoulos (FI) term, $`ϵ{\displaystyle \frac{g_s^2M_P^2}{192\pi ^2}}\mathrm{Tr}Q^{(A)}={\displaystyle \frac{g_s^2M_P^2}{192\pi ^2}}1392,`$ (5.9) generated in the $`U_A`$ $`D`$-term by the Green-Schwarz-Dine-Seiberg-Witten anomalous U(1) breaking mechanism . To see this, first note from Table C.3 that singlet states $`S_1^n`$ through $`S_6^n`$ carry the non-anomalous charge $`Q_1=3`$, while the remaining singlets have $`Q_1=0`$. Thus, $`D`$-flatness for $`U(1)_1`$ cannot be maintained if only singlets receive VEVs and one or more of the fields $`S_1^i`$ through $`S_6^i`$ are among those that do. Next, the singlets $`S_7`$ and $`\overline{S}_7`$ do not carry anomalous charge (only $`Q_4`$ and $`Q_5`$ charge), and so cannot help cancel the F-I term. The remaining singlets are simply $`S_8`$ and $`S_9`$, and their vector partners of opposite charges. $`S_8`$ carries anomalous charge $`Q_A=4`$ and non-zero non-anomalous charges $`Q_4=1`$, $`Q_5=4`$, $`Q_2^{}=6`$, $`Q_3^{}=42`$, while $`S_9`$ carries anomalous charge $`Q_A=4`$ and non-zero non-anomalous charges $`Q_4=1`$, $`Q_5=4`$, $`Q_2^{}=6`$, $`Q_3^{}=42`$. $`\overline{S}_8`$ and $`\overline{S}_9`$ carry respective opposite charges. Thus, we see that no combination of $`S_{7,8,9}`$ and $`\overline{S}_{7,8,9}`$ VEVs can simultaneously cancel the anomalous $`D_A`$-term contribution from the trace of $`U_A`$ and keep the D-terms for $`Q_2^{}`$ and $`Q_3^{}`$ flat. Therefore, some non-Abelian fields must take on VEVs in the process of cancelling the Tr$`Q_A`$ contribution to $`D_A`$ to maintain $`D`$-flatness. Of particular interest is whether $`SU(2)_i`$-charged fields take on VEVs in the parameter space of flat directions. Analysis of flat directions for these models is, however, beyond the scope of this letter. ## 6 Summary We have demonstrated how, under certain conditions, mirror symmetry is necessarily broken between the observable and hidden sector gauge groups of heterotic string models with mirror boundary conditions for observable and hidden sector worldsheet fermions. The observable/hidden sector gauge group mirror breaking occurs because of an unavoidable asymmetry in GSO projections. This effect can be induced in free fermionic models through an observable/hidden sector mirror pair of basis vectors, $`𝐚`$ and $`𝐚^{}`$, with the properties that: * Their vector sum yields new, independent gauge sectors (possibly after further basis vectors are added) $`𝐚+𝐚^{}+\mathrm{}`$. * They do not overlap with non-zero components in the observable and hidden sectors. * Their inner product is not equal $`0(\mathrm{mod}\mathrm{\hspace{0.17em}4})`$. Under these conditions the observable and hidden sector gauge states from some $`𝐚+𝐚^{}+\mathrm{}`$ sector (or sectors) will not be mirror images, since the observable gauge states surviving the $`k_{𝐚,𝐚+𝐚^{}+\mathrm{}}`$ GSO projections will be different from the hidden sector gauge states surviving the corresponding $`k_{𝐚^{},𝐚+𝐚^{}+\mathrm{}}`$ GSO projections. In the examples shown, starting with mirror Pati-Salam gauge groups $`[SU(4)_CSU(2)_LSU(2)_R]^O[SU(4)_CSU(2)_LSU(2)_R]^H`$, the observable sector $`SU(2)_{R(L)}`$ was transformed to the hidden sector by this necessary asymmetry of the GSO projections, enhancing the hidden sector gauge group to $`[SO(10)SU(2)_{R(L)}]^H`$. This transference of gauge rank from the observable sector to the hidden sector acts favorably for coupling strength renormalizations, allowing non-Abelian hidden sector coupling strengths to increase faster than observable sector coupling strengths, with decreasing energy below the string scale, leading to the formation of generally advantageous intermediate scale hidden sector condensates. ## 7 Acknowledgments This work is supported in part by the NASA/Texas Space Grant Consortium. Table A.1 THE NAHE SET | | $`\psi ^\mu `$ | $`\chi ^{12}`$ | $`\chi ^{34}`$ | $`\chi ^{56}`$ | $`\overline{\psi }^{1,\mathrm{},5}`$ | $`\overline{\eta }^1`$ | $`\overline{\eta }^2`$ | $`\overline{\eta }^3`$ | $`\overline{\psi ^{}}^{1,\mathrm{},5}`$ | $`\overline{\eta ^{}}^1`$ | $`\overline{\eta ^{}}^2`$ | $`\overline{\eta ^{}}^3`$ | | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | | 1 | 1 | 1 | 1 | 1 | 1,…,1 | 1 | 1 | 1 | 1,…,1 | 1 | 1 | 1 | | $`𝐒`$ | 1 | 1 | 1 | 1 | 0,…,0 | 0 | 0 | 0 | 0,…,0 | 0 | 0 | 0 | | $`𝐛_1`$ | 1 | 1 | 0 | 0 | 1,…,1 | 1 | 0 | 0 | 0,…,0 | 0 | 0 | 0 | | $`𝐛_2`$ | 1 | 0 | 1 | 0 | 1,…,1 | 0 | 1 | 0 | 0,…,0 | 0 | 0 | 0 | | $`𝐛_3`$ | 1 | 0 | 0 | 1 | 1,…,1 | 0 | 0 | 1 | 0,…,0 | 0 | 0 | 0 | NAHE SET b | | $`y^{3,\mathrm{},6}`$ | $`\overline{y}^{3,\mathrm{},6}`$ | $`y^{1,2},\omega ^{5,6}`$ | $`\overline{y}^{1,2},\overline{\omega }^{5,6}`$ | $`\omega ^{1,\mathrm{},4}`$ | $`\overline{\omega }^{1,\mathrm{},4}`$ | | --- | --- | --- | --- | --- | --- | --- | | 1 | 1,…,1 | 1,…,1 | 1,…,1 | 1,…,1 | 1,…,1 | 1,…,1 | | $`𝐒`$ | 0,…,0 | 0,…,0 | 0,…,0 | 0,…,0 | 0,…,0 | 0,…,0 | | $`𝐛_1`$ | 1,…,1 | 1,…,1 | 0,…,0 | 0,…,0 | 0,…,0 | 0,…,0 | | $`𝐛_2`$ | 0,…,0 | 0,…,0 | 1,…,1 | 1,…,1 | 0,…,0 | 0,…,0 | | $`𝐛_3`$ | 0,…,0 | 0,…,0 | 0,…,0 | 0,…,0 | 1,…,1 | 1,…,1 | Table A.2 Mirror Set | | $`\psi ^\mu `$ | $`\chi ^{12}`$ | $`\chi ^{34}`$ | $`\chi ^{56}`$ | $`\overline{\psi }^{1,\mathrm{},5}`$ | $`\overline{\eta }^1`$ | $`\overline{\eta }^2`$ | $`\overline{\eta }^3`$ | $`\overline{\psi ^{}}^{1,\mathrm{},5}`$ | $`\overline{\eta ^{}}^1`$ | $`\overline{\eta ^{}}^2`$ | $`\overline{\eta ^{}}^3`$ | | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | | $`𝐛_{}^{}{}_{1}{}^{}`$ | 1 | 1 | 0 | 0 | 0,…,0 | 0 | 0 | 0 | 1,…,1 | 1 | 0 | 0 | | $`𝐛_{}^{}{}_{2}{}^{}`$ | 1 | 0 | 1 | 0 | 0,…,0 | 0 | 0 | 0 | 1,…,1 | 0 | 1 | 0 | | $`𝐛_{}^{}{}_{3}{}^{}`$ | 1 | 0 | 0 | 1 | 0,…,0 | 0 | 0 | 0 | 1,…,1 | 0 | 0 | 1 | MIRROR SET b | | $`y^{1,\mathrm{},4}`$ | $`\overline{y}^{1,\mathrm{},4}`$ | $`y^{5,6},\omega ^{1,2}`$ | $`\overline{y}^{5,6},\overline{\omega }^{1,2}`$ | $`\omega ^{3,\mathrm{},6}`$ | $`\overline{\omega }^{3,\mathrm{},6}`$ | | --- | --- | --- | --- | --- | --- | --- | | $`𝐛_{}^{}{}_{1}{}^{}`$ | 1,…,1 | 1,…,1 | 0,…,0 | 0,…,0 | 0,…,0 | 0,…,0 | | $`𝐛_{}^{}{}_{2}{}^{}`$ | 0,…,0 | 0,…,0 | 1,…,1 | 1,…,1 | 0,…,0 | 0,…,0 | | $`𝐛_{}^{}{}_{3}{}^{}`$ | 0,…,0 | 0,…,0 | 0,…,0 | 0,…,0 | 1,…,1 | 1,…,1 | Table A.3 $`SO(10)SO(10)`$ Breaking and Generation Reduction | | $`\psi ^\mu `$ | $`\chi ^{12}`$ | $`\chi ^{34}`$ | $`\chi ^{56}`$ | $`\overline{\psi }^{1,\mathrm{},5}`$ | $`\overline{\eta }^1`$ | $`\overline{\eta }^2`$ | $`\overline{\eta }^3`$ | $`y,\overline{y},\omega \overline{,}\omega `$ | $`\overline{\psi ^{}}^{1,\mathrm{},5}`$ | $`\overline{\eta ^{}}^1`$ | $`\overline{\eta ^{}}^2`$ | $`\overline{\eta ^{}}^3`$ | | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | | $`𝐚`$ | 1 | 1 | 1 | 1 | 0,0,0,1,1 | 1 | 1 | 1 | $`\stackrel{}{0}`$ | 0,0,0,0,0 | 0 | 0 | 0 | | $`𝐚^{}`$ | 1 | 1 | 1 | 1 | 0,0,0,0,0 | 0 | 0 | 0 | $`\stackrel{}{0}`$ | 0,0,0,1,1 | 0 | 0 | 0 | Table B.1 Model 1 GSO Matrix $`𝐤`$ | $`k_{i,j}`$ | $`\mathrm{𝟏}`$ | $`𝐒`$ | $`𝐛_1`$ | $`𝐛_2`$ | $`𝐛_3`$ | $`𝐛_1^{}`$ | $`𝐛_2^{}`$ | $`𝐛_3^{}`$ | $`𝐚`$ | $`𝐚^{}`$ | | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | | $`\mathrm{𝟏}`$ | 0 | 0 | 1 | 1 | 1 | 1 | 1 | 1 | 0 | 0 | | $`𝐒`$ | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | | $`𝐛_1`$ | 1 | 1 | 1 | 1 | 1 | 0 | 0 | 0 | 0 | 1 | | $`𝐛_2`$ | 1 | 1 | 1 | 1 | 1 | 0 | 0 | 0 | 0 | 1 | | $`𝐛_3`$ | 1 | 1 | 1 | 1 | 1 | 0 | 0 | 0 | 0 | 1 | | $`𝐛_1^{}`$ | 1 | 1 | 0 | 0 | 0 | 1 | 1 | 1 | 1 | 0 | | $`𝐛_2^{}`$ | 1 | 1 | 0 | 0 | 0 | 1 | 1 | 1 | 1 | 0 | | $`𝐛_3^{}`$ | 1 | 1 | 0 | 0 | 0 | 1 | 1 | 1 | 1 | 0 | | $`𝐚`$ | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 1 | \*1 | | $`𝐚^{}`$ | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | \*0 | 1 | Table B.2 Model 1 Gauge Group | Observable | $`\overline{\psi }^1`$ | $`\overline{\psi }^2`$ | $`\overline{\psi }^3`$ | $`\overline{\psi }^3`$ | $`\overline{\psi }^5`$ | | --- | --- | --- | --- | --- | --- | | $`SU(4)_C`$ | 0 | 1 | -1 | 0 | 0 | | | 1 | 0 | -1 | 0 | 0 | | | 0 | 1 | 1 | 0 | 0 | | $`SU(2)_L`$ | 0 | 0 | 0 | 1 | -1 | | Shadow | $`\overline{\eta }^1`$ | $`\overline{\eta }^2`$ | $`\overline{\eta }^3`$ | $`\overline{y}_{1,3}`$ | $`\overline{y}_{2,4}`$ | $`\overline{y}_5\overline{w}_1`$ | $`\overline{y}_6\overline{\omega }_2`$ | $`\overline{\omega }_{3,5}`$ | $`\overline{\omega }_{4,6}`$ | $`\overline{\psi ^{}}^4`$ | $`\overline{\psi ^{}}^5`$ | $`\overline{\eta ^{}}^1`$ | $`\overline{\eta ^{}}^2`$ | $`\overline{\eta ^{}}^3`$ | | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | | $`SU(2)^3`$ | $`\frac{1}{2}`$ | -$`\frac{1}{2}`$ | 0 | 0 | $`\frac{1}{2}`$ | 0 | -$`\frac{1}{2}`$ | -$`\frac{1}{2}`$ | $`\frac{1}{2}`$ | 0 | 0 | -$`\frac{1}{2}`$ | $`\frac{1}{2}`$ | 0 | | | $`\frac{1}{2}`$ | 0 | -$`\frac{1}{2}`$ | $`\frac{1}{2}`$ | 0 | -$`\frac{1}{2}`$ | $`\frac{1}{2}`$ | 0 | -$`\frac{1}{2}`$ | 0 | 0 | -$`\frac{1}{2}`$ | 0 | $`\frac{1}{2}`$ | | | 0 | $`\frac{1}{2}`$ | -$`\frac{1}{2}`$ | -$`\frac{1}{2}`$ | $`\frac{1}{2}`$ | $`\frac{1}{2}`$ | 0 | -$`\frac{1}{2}`$ | 0 | 0 | 0 | 0 | -$`\frac{1}{2}`$ | $`\frac{1}{2}`$ | | $`SU(5)`$ | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 1 | 1 | 0 | 0 | 0 | | | $`\frac{1}{2}`$ | -$`\frac{1}{2}`$ | 0 | 0 | -$`\frac{1}{2}`$ | 0 | $`\frac{1}{2}`$ | -$`\frac{1}{2}`$ | $`\frac{1}{2}`$ | 0 | 0 | $`\frac{1}{2}`$ | -$`\frac{1}{2}`$ | 0 | | | 0 | $`\frac{1}{2}`$ | $`\frac{1}{2}`$ | 0 | 0 | 0 | -$`\frac{1}{2}`$ | 0 | -$`\frac{1}{2}`$ | $`\frac{1}{2}`$ | $`\frac{1}{2}`$ | 0 | $`\frac{1}{2}`$ | $`\frac{1}{2}`$ | | | 0 | $`\frac{1}{2}`$ | -$`\frac{1}{2}`$ | -$`\frac{1}{2}`$ | $`\frac{1}{2}`$ | -$`\frac{1}{2}`$ | 0 | $`\frac{1}{2}`$ | 0 | 0 | 0 | 0 | $`\frac{1}{2}`$ | -$`\frac{1}{2}`$ | | Hidden | $`\overline{\psi }^1`$ | $`\overline{\psi }^2`$ | $`\overline{\psi }^3`$ | $`\overline{\psi }^3`$ | $`\overline{\psi }^5`$ | $`\overline{\psi ^{}}^1`$ | $`\overline{\psi ^{}}^2`$ | $`\overline{\psi ^{}}^3`$ | $`\overline{\psi ^{}}^3`$ | $`\overline{\psi ^{}}^5`$ | | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | | $`SO(10)`$ | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 1 | -1 | | | 0 | 0 | 0 | $`\frac{1}{2}`$ | $`\frac{1}{2}`$ | 0 | 0 | -1 | $`\frac{1}{2}`$ | -$`\frac{1}{2}`$ | | | 0 | 0 | 0 | 0 | 0 | 0 | 1 | -1 | 0 | 0 | | | 0 | 0 | 0 | 0 | 0 | 1 | -1 | 0 | 0 | 0 | | | 0 | 0 | 0 | 0 | 0 | 1 | 1 | 0 | 0 | 0 | Table B.3 Model 1 Left-Handed States | $`n=1,2`$ | $`(4_C,2_L)_O`$ | $`(2^3,5)_S`$ | $`10_H`$ | $`4Q_1`$ | $`4Q_2`$ | $`4Q_3`$ | $`4Q_4`$ | $`4Q_5`$ | $`4Q_6`$ | | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | | Singlets | | | | | | | | | | | $`S_1`$ | (1,1) | (1,1,1,1) | (1) | 0 | -12 | -4 | -36 | 128 | 80 | | $`\overline{S}_1`$ | (1,1) | (1,1,1,1) | (1) | 0 | 12 | 4 | 36 | -128 | -80 | | $`S_2`$ | (1,1) | (1,1,1,1) | (1) | -4 | 4 | 4 | -4 | 176 | 0 | | $`\overline{S}_2`$ | (1,1) | (1,1,1,1) | (1) | 4 | -4 | -4 | 4 | -176 | 0 | | $`S_3`$ | (1,1) | (1,1,1,1) | (1) | -4 | 16 | 8 | 32 | 48 | -80 | | $`\overline{S}_3`$ | (1,1) | (1,1,1,1) | (1) | 4 | -16 | -8 | -32 | -48 | 80 | | Observable | | | | | | | | | | | $`QL_1^n`$ | (4,2) | (1,1,1,1) | (1) | 0 | -6 | -6 | 6 | -40 | -80 | | $`QL_2^n`$ | (4,2) | (1,1,1,1) | (1) | 0 | 0 | -4 | -56 | 0 | 0 | | $`QL_3^n`$ | (4,2) | (1,1,1,1) | (1) | -6 | 0 | 0 | 0 | 0 | 0 | | $`(q^cl^c)_1^n`$ | (-4,1) | (2,1,1,1) | (1) | -2 | -4 | 12 | -12 | 24 | -40 | | $`(q^cl^c)_2^n`$ | (-4,1) | (1,2,1,1) | (1) | 0 | -12 | 8 | -28 | 0 | 0 | | $`(q^cl^c)_3^n`$ | (-4,1) | (1,1,2,1) | (1) | 0 | -6 | 10 | -10 | -64 | -40 | | $`h_1^n`$ | (1,2) | (1,1,1,1) | (1) | 0 | -6 | -10 | -50 | -40 | -80 | | $`h_2^n`$ | (1,2) | (1,1,1,1) | (1) | -6 | -6 | -6 | 6 | -40 | -80 | | $`h_3^n`$ | (1,2) | (1,1,1,1) | (1) | -6 | 0 | -4 | -56 | 0 | 0 | | $`H_1^n`$ | (1,2) | (1,1,1,5) | (1) | 2 | -8 | 4 | 16 | 80 | -16 | | $`H_2^n`$ | (1,2) | (1,1,1,5) | (1) | 2 | 10 | 10 | -10 | -8 | -16 | | $`H_3^n`$ | (1,2) | (1,1,1,5) | (1) | 0 | -6 | 6 | 14 | -56 | 64 | Table B.3 Model 1 Left-Handed States, cont. | $`n=1,2`$ | $`(4_C,2_L)_O`$ | $`(2^3,5)_S`$ | $`10_H`$ | $`4Q_1`$ | $`4Q_2`$ | $`4Q_3`$ | $`4Q_4`$ | $`4Q_5`$ | $`4Q_6`$ | | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | | Shadow | | | | | | | | | | | $`F_1`$ | (1,1) | (1,1,1,5) | (1) | 0 | -12 | -4 | -36 | -96 | -16 | | $`\overline{F}_1`$ | (1,1) | (1,1,1,$`\overline{5}`$) | (1) | 0 | 12 | 4 | 36 | 96 | 16 | | $`F_2`$ | (1,1) | (1,1,1,5) | (1) | -4 | -8 | 0 | -40 | 80 | -16 | | $`\overline{F}_2`$ | (1,1) | (1,1,1,$`\overline{5}`$) | (1) | 4 | 8 | 0 | 40 | -80 | 16 | | $`F_3`$ | (1,1) | (1,1,1,5) | (1) | -4 | 4 | 4 | -4 | -48 | -96 | | $`\overline{F}_3`$ | (1,1) | (1,1,1,$`\overline{5}`$) | (1) | 4 | -4 | -4 | 4 | 48 | 96 | | $`\overline{G}_1^n`$ | (1,1) | (2,1,1,$`\overline{10}`$) | (1) | 0 | -6 | -2 | -18 | -48 | -8 | | $`\overline{G}_2^n`$ | (1,1) | (1,2,1,$`\overline{10}`$) | (1) | -2 | 2 | 2 | -2 | -24 | -48 | | $`\overline{G}_3^n`$ | (1,1) | (1,1,2,$`\overline{10}`$) | (1) | -2 | -4 | 0 | -20 | 40 | -8 | | $`X_1^n`$ | (1,1) | (2,1,1,1) | (1) | -2 | -4 | 16 | 44 | 24 | -40 | | $`X_2^n`$ | (1,1) | (2,1,1,1) | (1) | -2 | 2 | 18 | -18 | 64 | 40 | | $`X_3^n`$ | (1,1) | (2,1,1,1) | (1) | -2 | 2 | -14 | 14 | 112 | -40 | | $`X_4^n`$ | (1,1) | (2,1,1,1) | (1) | -2 | 20 | -8 | -12 | 24 | -40 | | $`Y_1^n`$ | (1,1) | (1,2,1,1) | (1) | 0 | -6 | -18 | -2 | 88 | 0 | | $`Y_2^n`$ | (1,1) | (1,2,1,1) | (1) | 6 | -12 | 8 | -28 | 0 | 0 | | $`Y_3^n`$ | (1,1) | (1,2,1,1) | (1) | 0 | -6 | 14 | -34 | 40 | 80 | | $`Y_4^n`$ | (1,1) | (1,2,1,1) | (1) | -2 | -4 | -16 | -4 | -48 | 80 | | $`Z_1^n`$ | (1,1) | (1,1,2,1) | (1) | -2 | 2 | -14 | 14 | -112 | 40 | | $`Z_2^n`$ | (1,1) | (1,1,2,1) | (1) | 0 | -6 | 14 | 46 | -64 | -40 | | $`Z_3^n`$ | (1,1) | (1,1,2,1) | (1) | 0 | 18 | -10 | -10 | -64 | -40 | | $`Z_4^n`$ | (1,1) | (1,1,2,1) | (1) | 6 | -6 | 10 | -10 | -64 | -40 | | $`U_1`$ | (1,1) | (2,2,1,1) | (1) | -4 | -8 | 0 | 40 | -24 | 40 | | $`\overline{U}_1`$ | (1,1) | (2,2,1,1) | (1) | 4 | 8 | 0 | -40 | 24 | -40 | | $`U_2`$ | (1,1) | (2,1,2,1) | (1) | -4 | 4 | 4 | -4 | -48 | 80 | | $`\overline{U}_2`$ | (1,1) | (2,1,2,1) | (1) | 4 | -4 | -4 | 4 | 48 | -80 | | $`U_3`$ | (1,1) | (1,2,2,1) | (1) | 0 | -12 | -4 | 44 | 24 | -40 | | $`\overline{U}_3`$ | (1,1) | (1,2,2,1) | (1) | 0 | 12 | 4 | -44 | -24 | 40 | | Hidden | | | | | | | | | | | $`T_1^n`$ | (1,1) | (1,1,1,1) | (16) | -2 | -16 | 0 | 0 | 0 | 0 | | $`T_2^n`$ | (1,1) | (1,1,1,1) | (16) | -2 | 2 | 6 | -26 | -88 | 0 | | $`T_3^n`$ | (1,1) | (1,1,1,1) | (16) | 0 | 0 | 4 | -24 | -24 | 0 | Table C.1 Model 2 GSO Matrix $`𝐤`$ | $`k_{i,j}`$ | $`\mathrm{𝟏}`$ | $`𝐒`$ | $`𝐛_1`$ | $`𝐛_2`$ | $`𝐛_3`$ | $`𝐛_1^{}`$ | $`𝐛_2^{}`$ | $`𝐛_3^{}`$ | $`𝐚`$ | $`𝐚^{}`$ | | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | | $`\mathrm{𝟏}`$ | 0 | 0 | 1 | 1 | 1 | 1 | 1 | 1 | 0 | 0 | | $`𝐒`$ | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | | $`𝐛_1`$ | 1 | 1 | 1 | 1 | 1 | 0 | 0 | 0 | 0 | 0 | | $`𝐛_2`$ | 1 | 1 | 1 | 1 | 1 | 0 | 0 | 0 | 0 | 0 | | $`𝐛_3`$ | 1 | 1 | 1 | 1 | 1 | 0 | 0 | 0 | 0 | 0 | | $`𝐛_1^{}`$ | 1 | 1 | 0 | 0 | 0 | 1 | 1 | 1 | 0 | 0 | | $`𝐛_2^{}`$ | 1 | 1 | 0 | 0 | 0 | 1 | 1 | 1 | 0 | 0 | | $`𝐛_3^{}`$ | 1 | 1 | 0 | 0 | 0 | 1 | 1 | 1 | 0 | 0 | | $`𝐚`$ | 0 | 0 | 1 | 1 | 1 | 0 | 0 | 0 | 1 | \*1 | | $`𝐚^{}`$ | 0 | 0 | 0 | 0 | 0 | 1 | 1 | 1 | \*0 | 1 | Table C.2 Model 2 Gauge Group | Observable | $`\overline{\psi }^1`$ | $`\overline{\psi }^2`$ | $`\overline{\psi }^3`$ | $`\overline{\psi }^3`$ | $`\overline{\psi }^5`$ | | --- | --- | --- | --- | --- | --- | | $`SU(4)_C`$ | 0 | 1 | -1 | 0 | 0 | | | 1 | 0 | -1 | 0 | 0 | | | 0 | 1 | 1 | 0 | 0 | | $`SU(2)_R`$ | 0 | 0 | 0 | 1 | 1 | | Shadow | $`\overline{\eta }^1`$ | $`\overline{\eta }^2`$ | $`\overline{\eta }^3`$ | $`\overline{y}_{1,3}`$ | $`\overline{y}_{2,4}`$ | $`\overline{y}_5\overline{w}_1`$ | $`\overline{y}_6\overline{\omega }_2`$ | $`\overline{\omega }_{3,5}`$ | $`\overline{\omega }_{4,6}`$ | $`\overline{\eta ^{}}^1`$ | $`\overline{\eta ^{}}^2`$ | $`\overline{\eta ^{}}^3`$ | | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | | $`SU(2)^3`$ | $`\frac{1}{2}`$ | $`\frac{1}{2}`$ | 0 | 0 | $`\frac{1}{2}`$ | 0 | $`\frac{1}{2}`$ | $`\frac{1}{2}`$ | $`\frac{1}{2}`$ | $`\frac{1}{2}`$ | $`\frac{1}{2}`$ | 0 | | | $`\frac{1}{2}`$ | 0 | $`\frac{1}{2}`$ | $`\frac{1}{2}`$ | 0 | $`\frac{1}{2}`$ | $`\frac{1}{2}`$ | 0 | $`\frac{1}{2}`$ | $`\frac{1}{2}`$ | 0 | $`\frac{1}{2}`$ | | | 0 | $`\frac{1}{2}`$ | $`\frac{1}{2}`$ | $`\frac{1}{2}`$ | $`\frac{1}{2}`$ | $`\frac{1}{2}`$ | 0 | $`\frac{1}{2}`$ | 0 | 0 | $`\frac{1}{2}`$ | $`\frac{1}{2}`$ | | $`SU(3)`$ | $`\frac{1}{2}`$ | $`\frac{1}{2}`$ | 0 | 0 | $`\frac{1}{2}`$ | 0 | $`\frac{1}{2}`$ | $`\frac{1}{2}`$ | $`\frac{1}{2}`$ | $`\frac{1}{2}`$ | $`\frac{1}{2}`$ | 0 | | | $`\frac{1}{2}`$ | 0 | $`\frac{1}{2}`$ | $`\frac{1}{2}`$ | 0 | $`\frac{1}{2}`$ | $`\frac{1}{2}`$ | 0 | $`\frac{1}{2}`$ | $`\frac{1}{2}`$ | 0 | $`\frac{1}{2}`$ | | Hidden | $`\overline{\psi }^1`$ | $`\overline{\psi }^2`$ | $`\overline{\psi }^3`$ | $`\overline{\psi }^3`$ | $`\overline{\psi }^5`$ | $`\overline{\psi ^{}}^1`$ | $`\overline{\psi ^{}}^2`$ | $`\overline{\psi ^{}}^3`$ | $`\overline{\psi ^{}}^3`$ | $`\overline{\psi ^{}}^5`$ | | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | | $`SO(10)`$ | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 1 | 1 | | | 0 | 0 | 0 | $`\frac{1}{2}`$ | $`\frac{1}{2}`$ | 0 | 0 | -1 | $`\frac{1}{2}`$ | $`\frac{1}{2}`$ | | | 0 | 0 | 0 | 0 | 0 | 0 | 1 | -1 | 0 | 0 | | | 0 | 0 | 0 | 0 | 0 | 1 | -1 | 0 | 0 | 0 | | | 0 | 0 | 0 | 0 | 0 | 1 | 1 | 0 | 0 | 0 | | $`SU(2)_L`$ | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 0 | 1 | -1 | Table C.3 Model 2 Left-Handed States | $`n=1,2`$ | $`(4_C,2_R)_O`$ | $`(2^3,3)^S`$ | $`(10,2_R)^H`$ | $`4Q_1`$ | $`4Q_2`$ | $`4Q_3`$ | $`4Q_4`$ | $`4Q_5`$ | $`4Q_6`$ | $`4Q_7`$ | | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | | Singlets | | | | | | | | | | | | $`S_1^n`$ | (1,1) | (1,1,1,1) | (1,1) | 12 | 0 | -4 | -8 | 16 | 0 | 0 | | $`S_2^n`$ | (1,1) | (1,1,1,1) | (1,1) | 12 | 0 | -4 | 8 | -16 | 0 | 0 | | $`S_3^n`$ | (1,1) | (1,1,1,1) | (1,1) | 12 | 0 | 0 | -6 | 0 | 4 | 24 | | $`S_4^n`$ | (1,1) | (1,1,1,1) | (1,1) | 12 | 0 | 0 | 6 | 0 | 4 | 24 | | $`S_5^n`$ | (1,1) | (1,1,1,1) | (1,1) | 12 | 0 | 0 | 2 | -16 | 4 | -24 | | $`S_6^n`$ | (1,1) | (1,1,1,1) | (1,1) | 12 | 0 | 0 | -2 | 16 | 4 | -24 | | $`S_7`$ | (1,1) | (1,1,1,1) | (1,1) | 0 | 0 | 0 | 8 | 32 | 0 | 0 | | $`\overline{S}_7`$ | (1,1) | (1,1,1,1) | (1,1) | 0 | 0 | 0 | -8 | -32 | 0 | 0 | | $`S_8`$ | (1,1) | (1,1,1,1) | (1,1) | 0 | 0 | 8 | 4 | 16 | 8 | 0 | | $`\overline{S}_8`$ | (1,1) | (1,1,1,1) | (1,1) | 0 | 0 | -8 | -4 | -16 | -8 | 0 | | $`S_9`$ | (1,1) | (1,1,1,1) | (1,1) | 0 | 0 | 8 | -4 | -16 | 8 | 0 | | $`\overline{S}_9`$ | (1,1) | (1,1,1,1) | (1,1) | 0 | 0 | -8 | 4 | 16 | -8 | 0 | | Observable | | | | | | | | | | | | $`QL_1^n`$ | (4,2) | (1,1,1,1) | (1,1) | -6 | -2 | -2 | 2 | -16 | 0 | -8 | | $`QL_2^n`$ | (4,2) | (1,1,1,1) | (1,1) | -6 | -2 | -2 | -2 | 16 | 0 | -8 | | $`QL_3^n`$ | (4,2) | (1,1,1,1) | (1,1) | -6 | -2 | 2 | 0 | 0 | 4 | 16 | | $`(q^cl^c)_1^n`$ | (-4,1) | (2,1,1,1) | (1,1) | 6 | -6 | -2 | 0 | 0 | -4 | 0 | | $`(q^cl^c)_2^n`$ | (-4,1) | (1,2,1,1) | (1,1) | 6 | -6 | 2 | -2 | -8 | 0 | 0 | | $`(q^cl^c)_3^n`$ | (-4,1) | (1,1,2,1) | (1,1) | 6 | -6 | 2 | 2 | 8 | 0 | 0 | | $`h_1`$ | (1,2) | (2,1,1,1) | (1,1) | 0 | -8 | 0 | 0 | 0 | 0 | 16 | | $`\overline{h}_1`$ | (1,2) | (2,1,1,1) | (1,1) | 0 | 8 | 0 | 0 | 0 | 0 | -16 | | $`h_2`$ | (1,2) | (1,2,1,1) | (1,1) | 0 | -8 | 0 | -4 | 8 | 0 | -8 | | $`\overline{h}_2`$ | (1,2) | (1,2,1,1) | (1,1) | 0 | 8 | 0 | 4 | -8 | 0 | 8 | | $`h_3`$ | (1,2) | (1,1,2,1) | (1,1) | 0 | -8 | 0 | 4 | -8 | 0 | -8 | | $`\overline{h}_3`$ | (1,2) | (1,1,2,1) | (1,1) | 0 | 8 | 0 | -4 | 8 | 0 | 8 | | $`H_1^n`$ | (1,2) | (2,1,1,3) | (1,1) | 4 | 0 | 4 | 0 | 0 | 0 | 0 | | $`H_2^n`$ | (1,2) | (1,2,1,3) | (1,1) | 4 | 0 | 0 | 2 | 8 | -4 | 0 | | $`H_3^n`$ | (1,2) | (1,1,2,3) | (1,1) | 4 | 0 | 0 | -2 | -8 | -4 | 0 | | $`B_1^n`$ | (6,1) | (1,1,1,1) | (1,1) | 0 | 4 | 4 | 6 | 0 | -4 | -8 | | $`B_2^n`$ | (6,1) | (1,1,1,1) | (1,1) | 0 | 4 | 0 | 0 | 0 | -8 | 16 | | $`B_3^n`$ | (6,1) | (1,1,1,1) | (1,1) | 0 | 4 | 4 | -6 | 0 | -4 | -8 | Table C.3 Model 2 Left-Handed States cont. | $`n=1,2`$ | $`(4_C,2_R)_O`$ | $`(2^3,3)^S`$ | $`(10,2_R)^H`$ | $`4Q_1`$ | $`4Q_2`$ | $`4Q_3`$ | $`4Q_4`$ | $`4Q_5`$ | $`4Q_6`$ | $`4Q_7`$ | | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | | Shadow | | | | | | | | | | | | $`F_1^n`$ | (1,1) | (1,1,1,$`\overline{3}`$) | (1,1) | -4 | -8 | -4 | 0 | 0 | 0 | 16 | | $`F_2^n`$ | (1,1) | (1,1,1,$`\overline{3}`$) | (1,1) | -4 | -8 | 0 | -6 | 0 | 4 | -8 | | $`F_3^n`$ | (1,1) | (1,1,1,$`\overline{3}`$) | (1,1) | -4 | -8 | 0 | 6 | 0 | 4 | -8 | | $`F_4^n`$ | (1,1) | (1,1,1,$`\overline{3}`$) | (1,1) | -4 | 8 | -4 | 0 | 0 | 0 | -16 | | $`F_5^n`$ | (1,1) | (1,1,1,$`\overline{3}`$) | (1,1) | -4 | 8 | 0 | -2 | 16 | 4 | 8 | | $`F_6^n`$ | (1,1) | (1,1,1,$`\overline{3}`$) | (1,1) | -4 | 8 | 0 | 2 | -16 | 4 | 8 | | $`F_7^n`$ | (1,1) | (1,1,1,3) | (1,1) | -8 | 0 | -8 | 0 | 0 | 0 | 0 | | $`\overline{F}_7^n`$ | (1,1) | (1,1,1,$`\overline{3}`$) | (1,1) | 8 | 0 | 8 | 0 | 0 | 0 | 0 | | $`F_8^n`$ | (1,1) | (1,1,1,3) | (1,1) | -8 | 0 | 0 | -4 | -16 | 8 | 0 | | $`\overline{F}_8^n`$ | (1,1) | (1,1,1,$`\overline{3}`$) | (1,1) | 8 | 0 | 0 | 4 | 16 | -8 | 0 | | $`F_9^n`$ | (1,1) | (1,1,1,3) | (1,1) | -8 | 0 | 0 | 4 | 16 | 8 | 0 | | $`\overline{F}_9^n`$ | (1,1) | (1,1,1,$`\overline{3}`$) | (1,1) | 8 | 0 | 0 | -4 | -16 | -8 | 0 | | $`X_1^n`$ | (1,1) | (2,2,1,1) | (1,1) | -12 | 0 | 0 | -2 | -8 | -4 | 0 | | $`X_2`$ | (1,1) | (2,2,1,1) | (1,1) | 0 | 0 | 0 | -4 | 8 | 0 | -24 | | $`X_3`$ | (1,1) | (2,2,1,1) | (1,1) | 0 | 0 | 0 | 4 | -8 | 0 | 24 | | $`Y_1^n`$ | (1,1) | (2,1,2,1) | (1,1) | -12 | 0 | 0 | 2 | 8 | -4 | 0 | | $`Y_2`$ | (1,1) | (2,1,2,1) | (1,1) | 0 | 0 | 0 | 4 | -8 | 0 | -24 | | $`Y_3`$ | (1,1) | (2,1,2,1) | (1,1) | 0 | 0 | 0 | -4 | 8 | 0 | 24 | | $`Z_1^n`$ | (1,1) | (1,2,2,1) | (1,1) | -12 | 0 | 4 | 0 | 0 | 0 | 0 | | $`Z_2`$ | (1,1) | (1,2,2,1) | (1,1) | 0 | 0 | 0 | -8 | 16 | 0 | 0 | | $`Z_3`$ | (1,1) | (1,2,2,1) | (1,1) | 0 | 0 | 0 | 8 | -16 | 0 | 0 | | Hidden | | | | | | | | | | | | $`T_1^n`$ | (1,1) | (1,1,1,1) | (10,1) | 0 | -4 | 4 | -2 | 16 | -4 | 8 | | $`T_2^n`$ | (1,1) | (1,1,1,1) | (10,1) | 0 | -4 | 4 | 2 | -16 | -4 | 8 | | $`T_3^n`$ | (1,1) | (1,1,1,1) | (10,1) | 0 | -4 | 0 | 0 | 0 | -8 | -16 | | $`\overline{F}_1^n`$ | (1,1) | (1,1,1,$`\overline{3}`$) | (1,2) | -4 | -4 | 4 | 0 | 0 | 0 | -16 | | $`\overline{F}_2^n`$ | (1,1) | (1,1,1,$`\overline{3}`$) | (1,2) | -4 | -4 | 0 | -6 | 0 | -4 | 8 | | $`\overline{F}_3^n`$ | (1,1) | (1,1,1,$`\overline{3}`$) | (1,2) | -4 | -4 | 0 | 6 | 0 | -4 | 8 | | $`D_1^n`$ | (1,1) | (1,1,1,1) | (1,2) | 12 | 4 | 4 | 0 | 0 | 0 | 16 | | $`D_2^n`$ | (1,1) | (1,1,1,1) | (1,2) | 12 | 4 | 0 | -2 | 16 | -4 | -8 | | $`D_3^n`$ | (1,1) | (1,1,1,1) | (1,2) | 12 | 4 | 0 | 2 | -16 | -4 | -8 | Table D.1 Hypercharge Components. | | $`\stackrel{~}{Q}_{BL}`$ | $`2T_3^R`$ | $`2T_3^L`$ | $`\stackrel{~}{Q}_{EM}`$ | $`=T_3^L+T_3^R+\frac{1}{2}\stackrel{~}{Q}_{BL}`$ | | --- | --- | --- | --- | --- | --- | | | $`\frac{1}{3}\stackrel{~}{Q}_C`$ | $`\frac{1}{2}\stackrel{~}{Q}_L`$ | | | $`=T_3^L+\frac{1}{2}Y`$ | | $`Q_L`$ | $`\frac{1}{3}`$ | $`0`$ | $`\pm 1`$ | | $`\frac{2}{3}`$, $`\frac{1}{3}`$ | | $`d_L^c`$ | $`\frac{1}{3}`$ | $`1`$ | $`0`$ | | $`\frac{1}{3}`$ | | $`u_L^c`$ | $`\frac{1}{3}`$ | $`1`$ | $`0`$ | | $`\frac{2}{3}`$ | | $`L_L`$ | $`1`$ | $`0`$ | $`\pm 1`$ | | $`0`$, $`1`$ | | $`e_L^c`$ | $`1`$ | $`1`$ | $`0`$ | | $`1`$ | | $`\nu _L^c`$ | $`1`$ | $`1`$ | $`0`$ | | $`0`$ |
warning/0506/quant-ph0506072.html
ar5iv
text
# Minimum error discrimination of Pauli channels ## 1 Introduction The concept of nonorthogonality of quantum states plays a relevant role in quantum computation and communication, cloning, and cryptography. Nonorthogonality is strongly related to the concept of distinguishability, and many measures have been introduced to compare quantum states and quantum processes . Since the seminal work of Helstrom on quantum hypothesis testing, the problem of discriminating nonorthogonal quantum states has received a lot of attention . Not very much work, however, has been devoted to the problem of discriminating general quantum operations, a part from the case of unitary transformations . Quantum operations describe any physically allowed transformation of quantum states, including unitary evolutions of closed systems and non unitary transformations of open quantum systems, such as systems interacting with a reservoir, or subjected to noise or measurements of any kind. The problem of discriminating quantum operations might be of great interest in quantum error correction , since knowing which error model is the proper one influences the choice of the coding strategy as well as the error estimation employed. Clearly, when a repeated use of the quantum operation is allowed, a full tomography can identify it. On the other hand, the minimum-error discrimination approach can be useful when a restricted number of uses of the quantum operation is considered, as in quantum hypothesis testing . In this paper we consider and solve the problem of discriminating with minimum error probability two given Pauli channels. Pauli channels represent the most general unital channels for qubits (e. g. bit flip, phase flip, depolarizing channels). Differently from the case of unitary transformations , we show that entanglement with an ancillary system at the input of the channel can strictly improve the discrimination. We prove that an arbitrary maximally entangled state is always an optimal input for the discrimination, and this holds true also for generalized Pauli channels in higher dimensional Hilbert spaces. However, the use of entanglement is not always needed to achieve optimality. In fact, we compare the strategies where either entangled or unentangled states are used at the input of the Pauli channels, and provide a necessary and sufficient condition in terms of the structure of the channels for which the ultimate minimum error probability can be achieved without the need of entanglement with an ancillary system. When such a condition is satisfied, the optimal input state is shown to be simply an eigenstate of one of the Pauli matrices. The paper is organized as follows. In Sec. II we briefly review the results for minimum error discrimination of two quantum states, and formulate the problem of discrimination of two quantum operations. In Sec. III we consider the problem for generalized Pauli channels in the scenario where entanglement with an ancillary system is allowed at the input of the channels. We prove that in any dimension an arbitrary maximally entangled state is always an optimal input for the discrimination, and the corresponding optimal measurement is a degenerate Bell measurement. In Sec. IV we find the optimal strategy for minimum error discrimination of two Pauli channels without entanglement assistance. Finally, in Sec. V we compare the two strategies and draw the conclusions. ## 2 Discriminating quantum operations In the problem of discrimination two quantum states $`\rho _1`$ and $`\rho _2`$, given with a priori probability $`p_1`$ and $`p_2=1p_1`$, respectively, one has to look for the two-values POVM $`\{P_i0,i=1,2\}`$ with $`P_1+P_2=I`$ that minimizes the error probability $`p_E(P_1,P_2)=p_1\text{Tr}[\rho _1P_2]+p_2\text{Tr}[\rho _2P_1].`$ (1) We can rewrite $`p_E(P_1,P_2)`$ $`=`$ $`p_1\text{Tr}[(p_1\rho _1p_2\rho _2)P_1]`$ (2) $`=`$ $`p_2+\text{Tr}[(p_1\rho _1p_2\rho _2)P_2]`$ $`=`$ $`{\displaystyle \frac{1}{2}}\left\{1\text{Tr}[(p_1\rho _1p_2\rho _2)(P_1P_2)]\right\},`$ where the third line can be obtained by summing and dividing the two lines above. The minimal error probability $`p_E\underset{P_1,P_2}{\mathrm{min}}p_E(P_1,P_2)`$ (3) can then be achieved by taking the orthogonal POVM $`\{P_1,P_2\}`$ made by the projectors on the support of the positive and negative part of the Hermitian operator $`p_1\rho _1p_2\rho _2`$, respectively, and hence one has $`p_E={\displaystyle \frac{1}{2}}\left(1p_1\rho _1p_2\rho _2_1\right),`$ (4) where $`A_1`$ denotes the trace norm of $`A`$. Equivalent expressions for the trace norm are the following $`A_1=\text{Tr}\sqrt{A^{}A}=\underset{U}{\mathrm{max}}|\text{Tr}[UA]|={\displaystyle \underset{n}{}}s_n(A),`$ (5) where the maximum is taken over all unitary operators, and $`\{s_n(A)\}`$ denote the singular values of $`A`$. In the case of Eq. (4), since the operator inside the norm is Hermitian, the singular values just corresponds to the absolute value of the eigenvalues. The problem of optimally discriminating two quantum operations $`_1`$ and $`_2`$ can be reformulated into the problem of finding in the input Hilbert space $``$ the state $`\rho `$ such that the error probability in the discrimination of the output states $`_1(\rho )`$ and $`_2(\rho )`$ is minimal. We are interested in the possibility of exploiting entanglement with an ancillary system in order to increase the distinguishability of the output states. In this case the output states to be discriminated will be of the form $`(_1_𝒦)\xi `$ and $`(_2_𝒦)\xi `$, where the input $`\xi `$ is generally a bipartite state of $`𝒦`$, and the quantum operations act just on the first party whereas the identity map $`=_𝒦`$ acts on the second. In the following we will denote with $`p_E^{}`$ the minimal error probability when a strategy without ancilla is adopted, and one has $`p_E^{}={\displaystyle \frac{1}{2}}\left(1\underset{\rho }{\mathrm{max}}p_1_1(\rho )p_2_2(\rho )_1\right).`$ (6) On the other hand, by allowing the use an ancillary system, we have $`p_E={\displaystyle \frac{1}{2}}\left(1\underset{\xi 𝒦}{\mathrm{max}}p_1(_1)\xi p_2(_2)\xi _1\right).`$ (7) The maximum of the trace norm in Eq. (7) is equivalent to the norm of complete boundedness of the map $`p_1_1p_2_2`$, and in fact for finite-dimensional Hilbert space one can just consider $`\text{dim}(𝒦)=\text{dim}()`$ . Moreover, from the linearity of quantum operations and the convexity of the trace norm , it follows that in both Eqs. (6) and (7) the maximum is achieved by pure states. Of course, $`p_Ep_E^{}`$. In the case of discrimination between two unitary transformations , one has $`p_E=p_E^{}`$, namely there is no need of entanglement with an ancillary system to achieve the ultimate minimum error probability. ## 3 Entanglement-assisted discrimination of generalized Pauli channels When the quantum operations can be realized from the same set of orthogonal unitaries as random unitary transformations , namely $`_i(\rho )={\displaystyle \underset{n}{}}q_n^{(i)}U_n\rho U_n^{},{\displaystyle \underset{n}{}}q_n^{(i)}=1,`$ (8) with $`\text{Tr}[U_m^{}U_n]=d\delta _{nm}`$, and the use of an ancillary system is allowed, one can evaluated the minimum error probability as follows. Let us define $`r_n=p_1q_n^{(1)}p_2q_n^{(2)}`$. One has $`\underset{|\psi 𝒦}{\mathrm{max}}p_1(_1)|\psi \psi |p_2(_2)|\psi \psi |_1`$ $`=\underset{|\psi 𝒦}{\mathrm{max}}{\displaystyle \underset{n}{}}r_n(U_nI)|\psi \psi |(U_n^{}I)_1`$ $`{\displaystyle \underset{n}{}}|r_n|\underset{|\psi 𝒦}{\mathrm{max}}(U_nI)|\psi \psi |(U_n^{}I)_1={\displaystyle \underset{n}{}}|r_n|.`$ (9) The bound is Eq. (9) can be saturated by the maximally entangled state $`|\mathrm{\Psi }={\displaystyle \frac{1}{\sqrt{d}}}{\displaystyle \underset{n=0}{\overset{d1}{}}}|n|n`$ (10) as we show in the following. Let us define $`A{\displaystyle \underset{n}{}}r_n(U_nI)|\mathrm{\Psi }\mathrm{\Psi }|(U_n^{}I)`$ (11) Notice that $`\mathrm{\Psi }|(U_n^{}I)(U_mI)|\mathrm{\Psi }={\displaystyle \frac{1}{d}}\mathrm{Tr}[U_m^{}U_n]=\delta _{nm},`$ (12) namely $`A`$ is diagonal on maximally entangled states. Then one has $`A_1=\mathrm{Tr}\sqrt{A^{}A}={\displaystyle \underset{n}{}}|r_n|.`$ (13) It follows that $`p_E={\displaystyle \frac{1}{2}}\left(1{\displaystyle \underset{n}{}}|r_n|\right).`$ (14) The corresponding measurement to be performed at the output of the channel is given by the projectors on support of the positive and negative part of the operator $`A`$ in Eq. (11), namely $`P_1={\displaystyle \underset{n_+}{}}(U_{n_+}I)|\mathrm{\Psi }\mathrm{\Psi }|(U_{n_+}^{}I),`$ $`P_2={\displaystyle \underset{n_{}}{}}(U_n_{}I)|\mathrm{\Psi }\mathrm{\Psi }|(U_n_{}^{}I),`$ (15) where the index $`n_+`$ ($`n_{}`$) are in correspondence with the positive (negative) elements of $`\{r_n\}`$. Notice that the set of projectors $`\{(U_nI)|\mathrm{\Psi }\mathrm{\Psi }|(U_n^{}I)\}`$ are orthogonal maximally entangled states, and hence the measurement is a degenerate Bell measurement . For the unitarily invariance property of the trace norm , the minimal error probability can always be achieved by using any arbitrary maximally entangled state at the input, namely $`(IV)|\mathrm{\Psi }`$ with $`V`$ unitary. The corresponding optimal measurement will be $`\{(IV)P_i(IV^{})\}`$, with $`\{P_i\}`$ as in Eq. (15). By dropping the condition of orthogonality of the $`\{U_n\}`$, one just obtains the bounds $`{\displaystyle \frac{1}{2}}\left(1{\displaystyle \underset{n}{}}|r_n|\right)p_E{\displaystyle \frac{1}{2}}\left(1A_1\right),`$ (16) since Eq. (9) gives the lower bound, whereas the upper bound is simply obtained by taking as input the maximally entangled state. ## 4 Discrimination of Pauli channels with no entanglement assistance In this section we consider the case of discrimination of two Pauli channels for qubits, namely $`^{(1)}(\rho )={\displaystyle \underset{n=0}{\overset{3}{}}}q_n^{(1)}\sigma _n\rho \sigma _n,^{(2)}(\rho )={\displaystyle \underset{n=0}{\overset{3}{}}}q_n^{(2)}\sigma _n\rho \sigma _n,`$ (17) where $`\{\sigma _0,\sigma _1,\sigma _2,\sigma _3\}=\{I,\sigma _x,\sigma _y,\sigma _z\}`$ and $`_{n=0}^3q_n^{(1)}=_{n=0}^3q_n^{(2)}=1`$. As shown in the previous section, the minimal error probability when entanglement with an ancillary system is used at the input is given by Eq. (14), where $`r_n=p_1q_n^{(1)}p_2q_n^{(2)},`$ (18) and $`p_1`$ and $`p_2`$ denote the a priori probabilities. Here, we are interested to understand when the entangled-input strategy is really needed to achieve the optimal discrimination, hence we derive in the following the optimal strategy with no ancillary system. According to Eq. (6) the minimal error probability is given by $`p_E^{}={\displaystyle \frac{1}{2}}\left(1\underset{|\psi }{\mathrm{max}}{\displaystyle \underset{n=0}{\overset{3}{}}}r_n\sigma _n|\psi \psi |\sigma _n_1\right)`$ (19) By parameterizing the pure state of the qubit as $`|\psi =\mathrm{cos}{\displaystyle \frac{\theta }{2}}|0+e^{i\varphi }\mathrm{sin}{\displaystyle \frac{\theta }{2}}|1,`$ (20) one has $`\xi {\displaystyle \underset{n=0}{\overset{3}{}}}r_n\sigma _n|\psi \psi |\sigma _n=`$ (21) $`\left(\begin{array}{cc}(r_0+r_3)\mathrm{cos}^2\frac{\theta }{2}+(r_1+r_2)\mathrm{sin}^2\frac{\theta }{2}& \frac{1}{2}\mathrm{sin}\theta [(r_0r_3)e^{i\varphi }+(r_1r_2)e^{i\varphi }]\\ \frac{1}{2}\mathrm{sin}\theta [(r_0r_3)e^{i\varphi }+(r_1r_2)e^{i\varphi }]& (r_0+r_3)\mathrm{sin}^2\frac{\theta }{2}+(r_1+r_2)\mathrm{cos}^2\frac{\theta }{2}\end{array}\right).`$ (24) The eigenvalues of $`\xi `$ are given by $`\lambda (\theta ,\varphi )_{1,2}={\displaystyle \frac{1}{2}}\{r_0+r_1+r_2+r_3\pm [\mathrm{cos}^2\theta [r_0+r_3r_1r_2]^2`$ $`+\mathrm{sin}^2\theta [(r_0r_3)^2+(r_1r_2)^2+2\mathrm{cos}(2\varphi )(r_0r_3)(r_1r_2)]]^{1/2}\}.`$ (25) We then have $`p_E^{}={\displaystyle \frac{1}{2}}\left(1\underset{\theta ,\varphi }{\mathrm{max}}[|\lambda _1(\theta ,\varphi )|+|\lambda _2(\theta ,\varphi )|]\right).`$ (26) Notice that the function $`f(\theta ,\varphi )(|\lambda _1(\theta ,\varphi )|+|\lambda _2(\theta ,\varphi )|)`$ can be rewritten as $`f(\theta ,\varphi )=`$ $`\mathrm{max}\{|r_0+r_1+r_2+r_3|,\{\mathrm{cos}^2\theta [r_0+r_3r_1r_2]^2`$ (27) $`+\mathrm{sin}^2\theta [(r_1r_2)^2+(r_0r_3)^2+2\mathrm{cos}(2\varphi )(r_0r_3)(r_1r_2)]\}^{1/2}\}.`$ The maximum over $`\theta ,\varphi `$ in Eq. (26) can be found just by comparing the values of $`f(\theta ,\varphi )`$ at the stationary points, namely $`\theta =k\pi `$, and $`\theta =\frac{(2k+1)\pi }{2},\varphi =l\frac{\pi }{2}`$, with $`k,l`$ integer. Since one has $`2[|\lambda _1(k\pi ,,\varphi )|+|\lambda _2(k\pi ,\varphi )|]`$ $`=|r_0+r_1+r_2+r_3+|r_0+r_3r_1r_2||+|r_0+r_1+r_2+r_3|r_0+r_3r_1r_2||`$ $`=2(|r_0+r_3|+|r_1+r_2|);`$ (28) $`2\left[\left|\lambda _1({\displaystyle \frac{(2k+1)\pi }{2}},l\pi )\right|+\left|\lambda _2({\displaystyle \frac{(2k+1)\pi }{2}},l\pi )\right|\right]`$ $`=|r_0+r_1+r_2+r_3+|r_0r_3+r_1r_2||+|r_0+r_1+r_2+r_3|r_0r_3+r_1r_2||`$ $`=2(|r_0+r_1|+|r_2+r_3|);`$ (29) $`2\left[\left|\lambda _1({\displaystyle \frac{(2k+1)\pi }{2}},{\displaystyle \frac{l\pi }{2}})\right|+\left|\lambda _2({\displaystyle \frac{(2k+1)\pi }{2}},{\displaystyle \frac{l\pi }{2}})\right|\right]`$ $`=|r_0+r_1+r_2+r_3+|r_0r_3r_1+r_2||+|r_0+r_1+r_2+r_3|r_0r_3r_1+r_2||`$ $`=2(|r_0+r_2|+|r_1+r_3|);`$ (30) one finally obtains $`p_E^{}={\displaystyle \frac{1}{2}}\left(1M\right),`$ (31) where $`M=\mathrm{max}\{|r_0+r_3|+|r_1+r_2|,|r_0+r_1|+|r_2+r_3|,|r_0+r_2|+|r_1+r_3|\}.`$ (32) The three cases inside the brackets corresponds to using an eigenstate of $`\sigma _z`$, $`\sigma _x`$, and $`\sigma _y`$, respectively, as input state of the unknown channel. The corresponding measurements to be performed at the output are the three Pauli matrices themselves. ## 5 Conclusion From comparing Eqs. (31) and (32) with Eq. (14), one can see that entanglement with an ancillary system is not needed to achieve the ultimate minimal error probability in the discrimination of the two Pauli channels of Eq. (17) as long as $`M=_{n=0}^3|r_n|`$, with $`r_n`$ given in Eq. (18). This happens when $`\mathrm{\Pi }_{n=0}^3r_n0`$. Hence, entanglement assistance is necessary if and only if all $`\{r_n\}`$ are different from zero, with three of them with the same sign, and the remaining one with the opposite sign. Among these cases, there are striking examples where the channels can be perfectly discriminated only by means of entanglement. This is the case of two channels of the form $`_1(\rho )={\displaystyle \underset{nm}{}}q_n\sigma _n\rho \sigma _n,_2(\rho )=\sigma _m\rho \sigma _m,`$ (33) with $`q_n0`$, and arbitrary a priori probability. This example can be simply understood, since the entanglement-assisted strategy increases the dimension of the Hilbert space such that the two possible output states will have orthogonal support. In conclusion, we considered the problem of discriminating two Pauli channels with minimal error probability. We showed that using maximally entangled states with an ancillary system at the input of the channel allows to achieve the optimal discrimination, and this holds true also for generalized Pauli channels in higher dimensional Hilbert spaces. In the case of qubits, we also found the minimal error probability for the discrimination strategy with no entanglement assistance, and showed that the optimal input states are the eigenstates of one of the Pauli matrices. By comparison, we then characterized in a simple way the instances where the optimal discrimination can be achieved without the need of entanglement. It could be interesting to look for similar conditions in the case of generalized Pauli channels in higher dimension. For this problem, one should translate the algebraic derivation in Sec. 4 into a more geometrical picture. This could result in better physical insights into why in some cases entanglement is not required to achieve minimum error discrimination. As in the field of state discrimination, one could also study the problem of optimal unambiguous discrimination of channels, where the unambiguity is paid by the possibility of getting inconclusive results from the measurement. Finally, an alternative approach is to consider the problem in the frequentistic scenario, instead of the Bayesian one, as it has been recently studied for state discrimination . In this case, one does not have a priori probabilities and has to maximize the worst probability of correct detection. This work has been sponsored by INFM through the project PRA-2002-CLON, and by EC and MIUR through the cosponsored ATESIT project IST-2000-29681 and Cofinanziamento 2003. ## References
warning/0506/cond-mat0506530.html
ar5iv
text
# Feshbach molecules in a one-dimensional Fermi gas ## I Introduction In a beautiful experiment Moritz *et al.* recently reported the observation of two-particle bound states of <sup>40</sup>K confined in a one-dimensional matter waveguide Moritz2005a . In the experiment an array of equivalent one-dimensional quantum system is realized by trapping a mixture of two hyperfine states of <sup>40</sup>K atoms in a two-dimensional optical lattice. The atoms are trapped at the intensity maxima and the radial confinement is only a fraction of the lattice period. At a given value of the magnetic field the binding energy $`E_B`$ of the bound states is probed by radio-frequency spectroscopy. Although Moritz *et al.* realized its limitations, the description of the experiment makes use of a single-channel model of radially confined atoms interacting with a pseudopotential Wilkens ; Bergeman2003a . Within this model the bound-state energy $`E_B`$ is related to the s-wave scattering length $`a`$ of the atoms by $`{\displaystyle \frac{a}{a_{}}}={\displaystyle \frac{\sqrt{2}}{\zeta (1/2,1/2E_B/2\mathrm{}\omega _{})}},`$ (1) where $`a_{}=\sqrt{\mathrm{}/m\omega _{}}`$, $`m`$ is the atomic mass, and $`\omega _{}`$ is the radial trapping frequency. To vary the scattering length, however, the experiment makes use of a Feshbach resonance at a magnetic field of $`B_0=202.1`$ Gauss. For such a Feshbach resonance a two-channel approach is physically more realistic. For the Feshbach problem the molecular binding energy $`E_B`$ always satisfies the equation review , $$E_B\delta (B)=\mathrm{}\mathrm{\Sigma }(E_B).$$ (2) Here the detuning $`\delta (B)=\mathrm{\Delta }\mu (BB_0)`$ varies as a function of the magnetic field and depends on the difference in magnetic moments $`\mathrm{\Delta }\mu `$ between the open and closed channels in the Feshbach problem. The resonance is located at the magnetic field strength $`B_0`$. For the homogeneous Fermi gas the molecular selfenergy is given by review $`\mathrm{}\mathrm{\Sigma }(E)=\left({\displaystyle \frac{g^2m^{3/2}}{4\pi \mathrm{}^3}}\right){\displaystyle \frac{i\sqrt{E}}{1i|a_{\mathrm{bg}}|\sqrt{mE/\mathrm{}^2}}},`$ (3) which leads to corrections to the single-channel result $`\mathrm{}^2/ma^2`$. Here $`g=\mathrm{}\sqrt{4\pi a_{\mathrm{bg}}\mathrm{\Delta }B\mathrm{\Delta }\mu /m}`$ is the atom-molecule coupling, $`\mathrm{\Delta }B`$ is the width of the Feshbach resonance, $`\mathrm{\Delta }\mu `$ is the difference in magnetic moments, and $`a_{\mathrm{bg}}`$ is the background scattering length. In Fig. 1 we show for this three-dimensional case the molecular binding energy for both the single and two-channel approaches, respectively. Whereas the single-channel results deviate significantly from the experimental data, there is an excellent agreement with the two-channel theory. It is therefore *a priori* not clear that in the one-dimensional case the single-channel theory as given by Eq. (1) is sufficiently accurate for the full range of magnetic fields explored by the experiment. In the following we derive the selfenergy for the confined case and make a comparisson with the experimental data. ## II Theory Two atoms in a waveguide near a Feshbach resonance are described by the following hamiltonian, $`H=H_\mathrm{a}+H_\mathrm{m}+V_{\mathrm{am}}.`$ (4) Here $`H_\mathrm{a}`$ represents the atomic contribution, $`H_\mathrm{m}`$ describes the bare molecules, and $`V_{\mathrm{am}}`$ is the atom-molecule coupling. Explicitely we have for the atoms, $`H_\mathrm{a}={\displaystyle \underset{i=1,2}{}}\left\{K_i+{\displaystyle \frac{m\omega _{}^2}{2}}(x_i^2+y_i^2)\right\}+V_{\mathrm{aa}}\delta (𝐫),`$ (5) with $`K_i=\mathrm{}^2_i^2/2m`$ the kinetic energy of atom $`i`$, $`V_{\mathrm{aa}}`$ is the strength of the nonresonant atom-atom interaction, and $`𝐫`$ the relative coordinate of the two atoms. The atoms are coupled to a molecular channel with a coupling $`V_{\mathrm{am}}`$. Near the resonance we have that $`V_{\mathrm{aa}}V_{\mathrm{am}}`$, which allows us to neglect the nonresonant atom-atom interaction in that case. For two atoms in the waveguide the two-channel Feshbach problem in the relative coordinate, after splitting off the center-of-mass motion, is then given by, $`\left(\begin{array}{cc}H_0& V_{\mathrm{am}}\\ V_{\mathrm{am}}& \delta _B\end{array}\right)\left(\begin{array}{c}|\psi _\mathrm{a}\\ |\psi _\mathrm{m}\end{array}\right)=E\left(\begin{array}{c}|\psi _\mathrm{a}\\ |\psi _\mathrm{m}\end{array}\right).`$ (6) Here the atomic Hamiltonian is $`H_0=\mathrm{}^2_𝐫^2/m+𝐫_{}^2/4`$, where $`_𝐫^2=_{}^2+_z^2`$ and $`𝐫_{}`$ is the radial component of $`𝐫`$. Only the relative part is relevant here, since only this part contains the interaction between the atoms. The bare detuning is denoted by $`\delta _B`$. The eigenstates $`|\psi _{n,k_z}`$ of $`H_0`$ that are relevant for an s-wave Feshbach resonance are a product state of a two-dimensional harmonic oscillator wave function in the radial direction and a plane wave along the axial direction. The associated energies are given by $`E_{n,k_z}=(2n+1)\mathrm{}\omega _{}+\mathrm{}^2k_z^2/m`$. The eigenstates of the two-dimensional harmonic oscillator that are relevant for s-wave scattering can be written as $`\psi _n(r_{},\varphi )=\left(2\pi a_{}^2\right)^{1/2}e^{r_{}^2/4a_{}^2}L_n^{(0)}(r_{}^2/2a_{}^2)`$, where $`L_n^{(0)}(x)`$ is the generalized Laguerre polynomial and $`\mathrm{}\omega _{}=\mathrm{}^2/ma_{}^2`$. From Eq. (6) we obtain the following equation determining the binding energy of the molecules: $`\psi _\mathrm{m}|V_{\mathrm{am}}{\displaystyle \frac{1}{EH_0}}V_{\mathrm{am}}|\psi _\mathrm{m}=E\delta _B.`$ (7) Using the above mentioned eigenstates of $`H_0`$, Eq. (7) can be written as $`{\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{dk_z}{2\pi }\frac{|\psi _\mathrm{m}|V_{\mathrm{am}}|\psi _{n,k_z}|^2}{EE_{n,k_z}}}=E\delta _B.`$ (8) Using also the usual pseudopotential approximation for the atom-molecule coupling, we have that $`𝐫|V_{\mathrm{am}}|\psi _\mathrm{m}=g\delta (𝐫)`$. Substituting this and performing the $`k_z`$ integration we obtain $`E\delta _B`$ $`=`$ $`\underset{r_{}0}{lim}{\displaystyle \frac{g^2m}{\sqrt{2}(4\pi a_{}\mathrm{}^2)}}`$ $`\times `$ $`{\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{e^{r_{}^2/4a_{}^2}L_n^{(0)}(r_{}^2/2a_{}^2)}{\sqrt{n+1/2E/2\mathrm{}\omega _{}}}}.`$ (9) The inverse square root $`1/\sqrt{n+1/2E/2\mathrm{}\omega _{}}`$ in the summand can be represented by the integral $`(2/\sqrt{\pi })_0^{\mathrm{}}𝑑te^{(n+1/2E/2\mathrm{}\omega _{})t^2}`$. To evaluate the sum over $`n`$ we substitute the above integral representation. The dependence on $`n`$ of the summand appears now in the exponent and in the degree of the Laguerre polynomial. As a result the sum can be directly evaluated by making use of the generating functions of the Laguerre polynomials, $$\underset{n=0}{\overset{\mathrm{}}{}}L_n^{(0)}(x)z^n=(1z)^1\mathrm{exp}\left(\frac{xz}{z1}\right).$$ (10) In our case we have $`z=e^{t^2}`$. Using this result and making the transformation $`y=t^2`$ we arrive at $`E\delta _B`$ $`=`$ $`\underset{r_{}0}{lim}{\displaystyle \frac{g^2m}{\sqrt{2\pi }(4\pi a_{}\mathrm{}^2)}}{\displaystyle _0^{\mathrm{}}}\mathrm{exp}\left({\displaystyle \frac{r_{}^2}{2a_{}^2}}{\displaystyle \frac{e^y}{e^y1}}\right)`$ $`\times `$ $`{\displaystyle \frac{\mathrm{exp}\left\{(1/2E/2\mathrm{}\omega _{})y\right\}}{\sqrt{y}(1e^y)}}dy`$ (11) For small values of $`y`$ the integrand in the above equation behaves as $`y^{3/2}e^{r^2/2y}`$. Note that we have $$\frac{1}{\sqrt{\pi }}_0^{\mathrm{}}𝑑yy^{3/2}e^{r^2/2y}=\sqrt{2}/r.$$ (12) We add and subtract this integral from Eq. (II) and in doing so we explicitely split off the $`1/r`$ divergence from the sum. The divergence in the selfenergy is energy independent and is related to the ultraviolet divergence that comes about because we have used pseudopotentials. To deal with this divergence we have to use the renormalized detuning instead of the bare detuning. The former is defined as $`\delta =\delta _Blim_{r0}mg^2/4\pi \mathrm{}^2r`$, where $`\delta =\mathrm{\Delta }\mu (BB_0)`$ is determined by the experimental value of the magnetic field $`B_0`$ at resonance and the magnetic moment difference $`\mathrm{\Delta }\mu =16/9`$ Bohr magneton for the <sup>40</sup>K atoms of interest. Note that, as expected, the required subtraction is exactly equal to the one needed in the absence of the optical lattice. In the latter case we have to subtract $`g^2𝑑𝐤m/\mathrm{}^2𝐤^2(2\pi )^3`$ review ; Holland , which can be interpreted as $`\delta =\delta _\mathrm{B}lim_{𝐫0}g^2𝑑𝐤e^{i𝐤𝐫}m/\mathrm{}^2𝐤^2(2\pi )^3`$. Using the renormalized detuning we find that the binding energy of the dressed molecules satisfies the desired equation $`E_B\delta (B)=\mathrm{}\mathrm{\Sigma }(E_B),`$ (13) where the molecular selfenergy for the harmonically confined one-dimensional system is given by $`\mathrm{}\mathrm{\Sigma }(E)={\displaystyle \frac{mg^2}{\sqrt{2}\left(4\pi a_{}\mathrm{}^2\right)}}\zeta (1/2,1/2E/2\mathrm{}\omega _{}).`$ (14) ## III Results and Discussion Using the selfenergy for the confined gas we can now solve for the binding energy in Eq. (13). The result is also shown in Fig. 1. We find an improved description of the experiment, although the differences with the single-channel prediction are small near resonance and only become large for larger detunings. This presents one way in which to experimentally probe these differences. Alternatively, it is also possible to directly measure the bare molecule fraction $`Z`$ of the dressed molecules Hulet , which is always equal to zero in the single-channel model. To be concrete we have for the dressed molecular wave function $$|\psi _{\mathrm{dressed}}=\sqrt{Z}|\psi _{\mathrm{closed}}+\sqrt{1Z}|\psi _{\mathrm{open}},$$ (15) where $`|\psi _{\mathrm{closed}}`$ is the wave function of the bare molecules and $`|\psi _{\mathrm{open}}`$ denotes the wave function of the atom pair in the open channel of the Feshbach resonance. With this application in mind we have plotted in Fig. 2 also the probability $`Z`$, which is determined from the selfenergy by $`Z=1/(1\mathrm{}\mathrm{\Sigma }(E_B)/E_B)`$. This work is supported by the Stichting voor Fundamenteel Onderzoek der Materie (FOM) and by the Nederlandse Organisatie voor Wetenschaplijk Onderzoek (NWO).
warning/0506/cs0506028.html
ar5iv
text
# Neyman-Pearson Detection of Gauss-Markov Signals in Noise: Closed-Form Error Exponent and Properties ## I Introduction In this paper, we consider the detection of correlated random signals using noisy observations $`y_i`$ under the Neyman-Pearson formulation. The null and alternative hypotheses are given by $$\begin{array}{ccc}H_0\hfill & :& y_i=w_i,i=1,2,\mathrm{},n,\hfill \\ H_1\hfill & :& y_i=s_i+w_i,\hfill \end{array}$$ (1) where $`\{w_i\}`$ is independent and identically distributed (i.i.d.) $`𝒩(0,\sigma ^2)`$ noise with a known variance $`\sigma ^2`$, and $`\{s_i\}`$ is the stochastic signal process correlated in time. We assume that $`\{s_i\}`$ is a Gauss-Markov process following a state-space model. An example of an application in which this type of problem arises is the detection of stochastic signals in large sensor networks, where it is reasonable to assume that signal samples taken at closely spaced locations are correlated, while the measurement noise is independent from sensor to sensor. In this paper, we are interested in the performance of the Neyman-Pearson detector for the hypotheses (1) with a fixed level (i.e., upper-bound constraint on the false-alarm probability) when the sample size $`n`$ is large. In many cases, the miss probability $`P_M`$ of the Neyman-Pearson detector with a fixed level decays exponentially as the sample size increases, and the error exponent is defined as the rate of exponential decay, i.e., $$K\stackrel{\mathrm{\Delta }}{=}\underset{n\mathrm{}}{lim}\frac{1}{n}\mathrm{log}P_M$$ (2) under the given false-alarm constraint. The error exponent is an important parameter since it gives an estimate of the number of samples required for a given detector performance; faster decay rate implies that fewer samples are needed for a given miss probability, or that better performance can be obtained with a given number of samples. Hence, the error exponent is a good performance index for detectors in the large sample regime. For the case of i.i.d. samples where each sample is drawn independently from the common null probability density $`p_0`$ or alternative density $`p_1`$, the error exponent under the fixed false-alarm constraint is given by the Kullback-Leibler information $`D(p_0||p_1)`$ between the two densities $`p_0`$ and $`p_1`$ (C. Stein ). For more general cases, the error exponent is given by the asymptotic Kullback-Leibler rate defined as the almost-sure limit of $$\frac{1}{n}\mathrm{log}\frac{p_{0,n}}{p_{1,n}}(y_1,\mathrm{},y_n)\text{as}n\mathrm{},$$ (3) under $`p_{0,n}`$, where $`p_{0,n}`$ and $`p_{1,n}`$ are the null and alternative joint densities of $`y_1,\mathrm{},y_n`$, respectively, assuming that the limit existsErgodic cases are examples for which this limit exists.. However, the closed-form calculation of (3) is available only for restricted cases. One such example is the discrimination between two autoregressive (AR) signals with distinct parameters under the two hypotheses . In this case, the joint density, $`p_{j,n}`$, is easily decomposed using the Markov property under each hypothesis, and the calculation of the rate is straightforward. However, for the problem of (1) this approach is not available since the observation samples under the alternative hypothesis do not possess the Markov property due to the additive noise, even if the signal itself is Markovian; i.e., the alternative is a hidden Markov model. ### I-A Summary of Results Our approach to this problem is to exploit the state-space model. The state-space approach in detection is well established in calculation of the log-likelihood ratio (LLR) for correlated signals. With the state-space model, the LLR is expressed through the innovations representation and the innovations are easily obtained by the Kalman filter. The key idea for the closed-form calculation of the error exponent for the hidden Markov model is based on the properties of innovations. Since the innovations process is independent from time to time, the joint density under $`H_1`$ is given by the product of marginal densities of the innovations, and the LLR is given by a function of the sum of squares of the innovations; this functional form facilitates the closed-form calculation of (3). By applying this state-space approach, we derive a closed-form expression for the error exponent $`K`$ for the miss probability of the Neyman-Pearson detector for (1) of fixed false-alarm probability, $`P_F=\alpha `$. We next investigate the properties of the error exponent using the obtained closed-form expression. We explore the asymptotic relationship between the innovations approach and the spectrum of the observation. We show that the error exponent $`K`$ is a function of the signal-to-noise ratio (SNR) and the correlation, and has different behavior with respect to (w.r.t.) the correlation strength depending on the SNR. We show a sharp phase transition at SNR = 1: at high SNR, $`K`$ is monotonically decreasing as a function of the correlation, while at low SNR, on the other hand, there exists an optimal correlation value that yields the maximal $`K`$. We also make a connection between the asymptotic behavior of the Kalman filter and that of the Neyman-Pearson detector. It is shown that the error exponent is determined by the asymptotic (or steady-state) variances of the innovations under $`H_0`$ and $`H_1`$ together with the noise variance. ### I-B Related Works The detection of Gauss-Markov processes in Gaussian noise is a classical problem. See and references therein. Our work focuses on the performance analysis as measured by the error exponent, and relies on the connection between the likelihood ratio and the innovations process as described by Schweppe . In addition to the calculation of the LLR, the state-space approach has also been used in the performance analysis in this detection problem. Exploiting the state-space model, Schweppe obtained a differential equation for the Bhattacharyya distance between two Gaussian processes, which gives an upper bound on the average error probability under a Bayesian formulation. There is an extensive literature on the large deviations approach to the analysis of the detection of Gauss-Markov processes . Many of these results rely on the extension of Cramer’s theorem by Gärtner and Ellis and the properties of the asymptotic eigenvalue distributions of Toeplitz matrices. To find the rate function, however, this approach usually requires an optimization that requires nontrivial numerical methods except in some simple cases, and the rate is given as an integral of the spectrum of the observation process; closed-form expressions are difficult to obtain except for the case of a noiseless autoregressive (AR) process in discrete-time and its continuous-time counterpart, the Ornstein-Uhlenbeck process. In addition, most results have been obtained for a fixed threshold for the normalized LLR test, which results in expressions for the rate as a function of the threshold. For ergodic cases, however, the normalized LLR converges to a constant under the null hypothesis and the false alarm probability also decays exponentially for a fixed threshold. Hence, a detector with a fixed threshold is not optimal in the Neyman-Pearson sense since it does not use the false-alarm constraint fully; i.e., the optimal threshold is a function of sample size. ### I-C Notation and Organization We will make use of standard notational conventions. Vectors and matrices are written in boldface with matrices in capitals. All vectors are column vectors. For a scalar $`z`$, $`z^{}`$ denotes the complex conjugate. For a matrix $`𝐀`$, $`𝐀^T`$ and $`𝐀^H`$ indicate the transpose and Hermitian transpose, respectively. $`\text{det}(𝐀)`$ and $`\text{tr}(𝐀)`$ denote the determinant and trace of $`𝐀`$, respectively. $`𝐀(l,m)`$ denotes the element of the $`l`$th row and $`m`$th column, and $`\{\lambda _k(𝐀)\}`$ denotes the set of all eigenvalues of $`𝐀`$. We reserve $`𝐈_m`$ for the identity matrix of size $`m`$ (the subscript is included only when necessary). For a sequence of random vectors $`𝐱_n`$, $`𝔼_j\{𝐱_n\}`$ is the expectation of $`𝐱_n`$ under probability density $`p_{j,n},j=0,1`$. The notation $`𝐱𝒩(𝝁,𝚺)`$ means that $`𝐱`$ has the multivariate Gaussian distribution with mean $`𝝁`$ and covariance $`𝚺`$. The paper is organized as follows. The data model is described in Section II. In Section III, the closed-form error exponent is obtained via the innovations approach representation. The properties of the error exponent are investigated in Section IV, and the extension to the vector case is provided in Section V. Simulation results are presented to demonstrate the predicted behavior in Section VI, followed by the conclusion in Section VII. ## II Data Model For the purposes of exposition, we will focus primarily on the case in which the signal is generated by a scalar time-invariant state space model. The more general vector case will be considered below. In particular, we assume that the signal process $`\{s_i\}`$ has a time-invariant state-space structure $`s_{i+1}`$ $`=`$ $`as_i+u_i,i=1,\mathrm{},n,`$ (4) $`s_1`$ $``$ $`𝒩(0,\mathrm{\Pi }_0),`$ $`u_i`$ $`\stackrel{i.i.d.}{}`$ $`𝒩(0,Q),Q=\mathrm{\Pi }_0(1a^2),`$ where $`a`$ and $`\mathrm{\Pi }_0`$ are known scalars with $`0a1`$ and $`\mathrm{\Pi }_00`$. We assume that the process noise $`\{u_i\}`$ is independent of the measurement noise $`\{w_i\}`$ and the initial state $`s_1`$ is independent of $`u_i`$ for all $`i`$. Notice that the signal sequence $`\{s_i\}`$ forms a stationary process for this choice of $`Q`$. Due to this stationarity, the signal variance is $`\mathrm{\Pi }_0`$ for all $`i`$, and the SNR $`\mathrm{\Gamma }`$ for the observations is thus given by $$\mathrm{\Gamma }=\frac{\mathrm{\Pi }_0}{\sigma ^2}.$$ (5) Notice that the value of $`a`$ determines the amount of correlation between signal samples. For an i.i.d. signal we have $`a=0`$ and all the signal power results from the process noise $`\{u_i\}`$. When the signal is perfectly correlated on the other hand, $`a=1`$ and the signal depends only on the realization of the initial state $`s_1`$. The autocovariance function $`r_s()`$ of the signal process $`\{s_i\}`$ is given by $$r_s(ij)\stackrel{\mathrm{\Delta }}{=}𝔼\{s_is_j\}=\mathrm{\Pi }_0a^{|ij|}.$$ (6) As seen in (1), the observation $`y_i`$ under the alternative hypothesis is given by a sum of signal sample $`s_i`$ and independent noise $`w_i`$. Thus, the observation sequence $`\{y_i\}`$ under $`H_1`$ is not a Markov process due to the presence of the additive noise even if the signal is Markovian. Let $`r_y^{(j)}()`$ denote the autocovariance function of the observation process $`\{y_i\}`$ under $`H_j`$, i.e., $$r_y^{(j)}(mn)=𝔼_j\{y_my_n\},$$ (7) and let $`S_y^{(j)}(\omega )`$ be the spectrum of the observation process under $`H_j`$, i.e., $$S_y^{(j)}(\omega )=\underset{k=\mathrm{}}{\overset{\mathrm{}}{}}r_y^{(j)}(k)e^{jk\omega },\pi \omega \pi .$$ (8) Then, the spectra of the observation process under $`H_0`$ and $`H_1`$ are given by $$S_y^{(0)}(\omega )=\sigma ^2,S_y^{(1)}(\omega )=\sigma ^2+S_s(\omega ),\pi \omega \pi ,$$ (9) where the signal spectrum under the state-space model is given by the Poisson kernel: $$S_s(\omega )=\frac{\mathrm{\Pi }_0(1a^2)}{12a\mathrm{cos}\omega +a^2},0a<1.$$ (10) ## III Error Exponent for Gauss-Markov Signal in Noise In this section, we derive the error exponent of the Neyman-Pearson detector with a fixed level $`\alpha (0,1)`$ for the Gauss-Markov signal described by (4) embedded in noisy observations. A general approach to the error exponent of Neyman-Pearson detection of Gaussian signals can be framed in the spectral domain. It is well known that the Kullback-Leibler information between two zero-mean Gaussian distributions $`p_0=𝒩(0,\sigma _0^2)`$ and $`p_1=𝒩(0,\sigma _1^2)`$ is given by $$D(p_0||p_1)=\frac{1}{2}\mathrm{log}\frac{\sigma _1^2}{\sigma _0^2}+\frac{1}{2}\frac{\sigma _0^2}{\sigma _1^2}\frac{1}{2}.$$ (11) As noted above, this quantity gives the error exponent in the case of an i.i.d. Gaussian signal. In more general cases with correlated Gaussian signals, the error exponent can similarly be obtained using the asymptotic properties of covariance matrices. Let $`𝐲_n`$ be the random vector of observation samples $`y_i`$ defined as $$𝐲_n\stackrel{\mathrm{\Delta }}{=}[y_1,y_2,\mathrm{},y_n]^T.$$ (12) For two distributions $`p_{0,n}(𝐲_n)=𝒩(\mathrm{𝟎},𝚺_{0,n})`$ and $`p_{1,n}(𝐲_n)=𝒩(\mathrm{𝟎},𝚺_{1,n})`$, the error exponent is given by the almost-sure limit of the Kullback-Leibler rate: $$\frac{1}{n}\mathrm{log}\frac{p_{0,n}}{p_{1,n}}(𝐲_n)=\frac{1}{n}\left(\frac{1}{2}\mathrm{log}\left(\frac{det(𝚺_{1,n})}{det(𝚺_{0,n})}\right)+\frac{1}{2}𝐲_n^T(𝚺_{1,n}^1𝚺_{0,n}^1)𝐲_n\right),$$ (13) under $`p_{0,n}`$ . Using the asymptotic distribution of the eigenvalues of a Toeplitz matrix, we have $$\underset{n\mathrm{}}{lim}\frac{1}{n}\mathrm{log}(det(𝚺_{j,n}))=\frac{1}{2\pi }_0^{2\pi }\mathrm{log}S_y^{(j)}(\omega )𝑑\omega ,j=0,1,$$ (14) where $`S_y^{(j)}(\omega )`$ is the spectrum of $`\{y_i\}`$ which is assumed to have finite lower and upper bounds under distribution $`𝐲_np_{j,n}`$. The limiting behavior of $`n^1𝐲_n^T𝚺_{j,n}^1𝐲_n`$ is also known and is given by (assuming that the true distribution of $`\{y_i\}`$ is $`p_{0,n}`$) $`\underset{n\mathrm{}}{lim}{\displaystyle \frac{1}{n}}𝐲_n^T𝚺_{1,n}^1𝐲_n`$ $`=`$ $`{\displaystyle \frac{1}{2\pi }}{\displaystyle _0^{2\pi }}{\displaystyle \frac{S_y^{(0)}(\omega )}{S_y^{(1)}(\omega )}}𝑑\omega ,`$ (15) $`\underset{n\mathrm{}}{lim}{\displaystyle \frac{1}{n}}𝐲_n^T𝚺_{0,n}^1𝐲_n`$ $`=`$ $`1,`$ (16) where the limit is in the almost-sure sense convergence under $`H_0`$, provided that $`S_y^{(0)}(\omega )`$ and $`S_y^{(1)}(\omega )`$ are continuous and strictly positive. (See Lemma 1 and 2 in and Prop. 10.8.2 and 10.8.3 in .) Combining (13)-(16), the error exponent for two zero-mean stationary Gaussian processes is thus given by $`K`$ $`=`$ $`\underset{n\mathrm{}}{lim}{\displaystyle \frac{1}{n}}\mathrm{log}{\displaystyle \frac{p_{0,n}}{p_{1,n}}}(𝐲_n)\text{a.s.}[p_{0,n}],`$ (17) $`=`$ $`{\displaystyle \frac{1}{2\pi }}{\displaystyle _0^{2\pi }}\left({\displaystyle \frac{1}{2}}\mathrm{log}{\displaystyle \frac{S_y^{(1)}(\omega )}{S_y^{(0)}(\omega )}}+{\displaystyle \frac{S_y^{(0)}(\omega )}{2S_y^{(1)}(\omega )}}{\displaystyle \frac{1}{2}}\right)𝑑\omega ,`$ (18) $`=`$ $`{\displaystyle \frac{1}{2\pi }}{\displaystyle _0^{2\pi }}D(𝒩(0,S_y^{(0)}(\omega ))||𝒩(0,S_y^{(1)}(\omega )))d\omega .`$ (19) Intuitively, the error exponent (19) can be explained from (11) using the frequency binning argument used to obtain the channel capacity of Gaussian channel with colored noise from that of independent parallel Gaussian channels . The spectral form (19) of the error exponent is valid for a wide class of stationary Gaussian processes including the autoregressive moving average (ARMA) processes and the hidden Markov model (1) - (4). For the detection (1) under the scalar state space model (4), we have $$𝚺_{0,n}=𝐈,𝚺_{1,n}=𝚺_{s,n}+𝐈,$$ (20) where $`𝚺_{s,n}(l,m)=\mathrm{\Pi }_0a^{|lm|},l,m=1,\mathrm{},n`$, and two spectra under $`H_0`$ and $`H_1`$ are given by (9). However, it is not straightforward to obtain a closed-form expression for (19) except in some special cases, e.g., when both of the two distributions of $`\{y_i\}`$ under $`H_0`$ and $`H_1`$ have the Markov property . In the remainder of the paper, we focus on the derivation of a closed-form expression for the error exponent $`K`$ of the miss probability for (1) - (4) by exploiting the state-space structure under the alternative hypothesis. We do so by making a connection with Kalman filtering. Our expressions will allow us to investigate the properties of the error exponent. ### III-A Closed-Form Error Exponent via Innovations Approach ###### Theorem 1 (Error exponent) For the Neyman-Pearson detector for the hypotheses (1) with level $`\alpha (0,1)`$ (i.e. $`P_F\alpha `$) and $`0a1`$, the error exponent of the miss probability is given by $`K`$ $`=`$ $`{\displaystyle \frac{1}{2}}\mathrm{log}{\displaystyle \frac{R_e}{\sigma ^2}}+{\displaystyle \frac{1}{2}}{\displaystyle \frac{\stackrel{~}{R}_e}{R_e}}{\displaystyle \frac{1}{2}},`$ (21) independently of the value of $`\alpha `$, where $`R_e`$ and $`\stackrel{~}{R}_e`$ are the steady-state variances of the innovations process of $`\{y_i\}`$ calculated under $`H_1`$ and $`H_0`$, respectively. Specifically, $`R_e`$ and $`\stackrel{~}{R}_e`$ are given by $$R_e=P+\sigma ^2,$$ (22) and $$\stackrel{~}{R}_e=\sigma ^2\left(1+\frac{a^2P^2}{P^2+2\sigma ^2P+(1a^2)\sigma ^4}\right),$$ (23) where $$P=\frac{1}{2}\sqrt{[\sigma ^2(1a^2)Q]^2+4\sigma ^2Q}\frac{1}{2}\sigma ^2(1a^2)+\frac{Q}{2}.$$ (24) Here, $`P`$ is the steady-state error variance of the minimum mean square error (MMSE) estimator for the signal $`s_i`$ under the model $`H_1`$. Note that the error exponent (21) is thus a closed-form of (19) for the state-space model. Proof: See the Appendix. Theorem 1 follows from the fact that the almost-sure limit (3) of the normalized log-likelihood ratio under $`H_0`$ is the error exponent for general ergodic cases. To make the closed-form calculation of the error exponent tractable for the hidden Markov structure of $`\{y_i\}`$, we express the log-likelihood ratio through the innovations representation ; the log-likelihood ratio is given by a function of the sum of squares of the innovations on which the strong law of large numbers (SLLN) is applied. The calculated innovations are true in the sense that they form an independent sequence only under $`H_1`$, i.e., when the signal actually comes from the state-space model. It is worth noting that $`\stackrel{~}{R}_e`$ is the steady-state variance of the “innovations” calculated as if the observations result from the alternative, but are actually from the null hypothesis. In this case, the “innovation” sequence becomes the output of a recursive (whitening) filter driven by an i.i.d. process $`\{y_i\}`$ since the Kalman filter converges to the recursive Wiener filter for time-invariant stable systems . The relationship between the spectral-domain approach and the innovations approach is explained by the canonical spectral factorization, which is well established for the state-space model. The asymptotic variance of the innovations sequence is the key parameter in both cases. The relationship between the asymptotic performance of the Neyman-Pearson detector and that of the Kalman filter is evident in (21) for the state-space model. In both cases, the innovations process plays a critical role in characterizing the performance, and the asymptotic variance of the innovation is sufficient for the calculation of the error exponent for the Neyman-Pearson detector and the steady-state error variance for the Kalman filter. ## IV Properties of Error Exponent In this section, we investigate the properties of the error exponent derived in the preceding section. We particularly examine the large sample error behavior with respect to the correlation strength and SNR. We show that the intensity of the additive noise significantly changes the error behavior with respect to the correlation strength, and the error exponent has a distinct phase transition in behavior with respect to the correlation strength depending on SNR. ###### Theorem 2 ($`K`$ vs. correlation) The error exponent $`K`$ is a continuous function of the correlation coefficient $`a`$ ($`0a1`$) for a given SNR $`0`$. The error exponent as a function of correlation strength is characterized by the following: * For SNR $`\mathrm{\Gamma }1`$, $`K`$ is monotonically decreasing as the correlation strength increases (i.e. $`a1`$); * For SNR $`\mathrm{\Gamma }<1`$, there exists a non-zero value $`a^{}`$ of the correlation coefficient that achieves the maximal $`K`$, and $`a^{}`$ is given by the solution of the following equation. $$[1+a^2+\mathrm{\Gamma }(1a^2)]^22\left(r_e+\frac{a^4}{r_e}\right)=0,$$ (25) where $`r_e=R_e/\sigma ^2`$. Furthermore, $`a^{}`$ converges to one as $`\mathrm{\Gamma }`$ goes to zero. Proof: See the Appendix. We first note that Theorem 2 shows that an i.i.d. signal gives the best error performance for a given SNR $`1`$ with the maximal error exponent being $`D(𝒩(0,\sigma ^2)||𝒩(0,\mathrm{\Pi }_0+\sigma ^2))`$. (In this case, Theorem 1 reduces to Stein’s lemma.) The intuition behind this result is that the signal component in the observation is strong at high SNR, and the innovations (the new information) provide more benefit to the detector than the noise averaging effect present for correlated observations. That is, simple radiometry provides sufficient detection power when the signal level is above that of the noise. Fig. 1 shows the error exponent as a function of the correlation coefficient $`a`$ for SNR $`\mathrm{\Gamma }=`$ 10 dB. The monotonicity of the error exponent is clearly seen; moreover, we see that the amount of decrease becomes larger as $`a`$ increases. Notice also that the amount of performance degradation from the i.i.d. case is not severe for weak correlation and the error exponent decreases suddenly near $`a=1`$ and eventually becomes zero at $`a=1`$. (It is easy to show that the miss probability decays with $`\mathrm{\Theta }(\frac{1}{\sqrt{n}})`$ for any SNR at $`a=1`$.) In contrast, the error exponent does not decrease monotonically in $`a`$ for SNR $`<1`$, and there exists an optimal correlation as shown in Fig. 2. It is seen that the i.i.d. case no longer gives the error performance for a fixed SNR. The error exponent initially increases as $`a`$ increases, and then decreases to zero as $`a`$ approaches one. As the SNR further decreases (see the cases of -6 dB and -9dB) the error exponent decreases for a fixed correlation strength, and the value of $`a`$ achieving the maximal error exponent is shifted closer to one. At low SNR the noise in the observation dominates. So, intuitively, making the signal more correlated provides greater benefit of noise averaging. The lower the SNR, the stronger we would like the correlation to be in order to compensate for the dominant noise power, as shown in Fig. 2. However, excessive correlation in the signal does not provide new information by observation, and the error exponent ultimately converges to zero as $`a`$ approaches one. Notice that the ratio of the error exponent for the optimal correlation to that for the i.i.d. case becomes large as SNR decreases. Hence, the improvement due to optimal correlation can be large for low SNR cases. Fig. 3 shows the value of $`a`$ that maximizes the error exponent as a function of SNR. As shown in the figure, unit SNR is a transition point between two different behavioral regimes of the error exponent with respect to correlation strength, and the transition is very sharp; the optimal correlation strength $`a`$ approaches one rapidly once SNR becomes smaller than one. The behavior of the error exponent with respect to SNR is given by the following theorem. ###### Theorem 3 ($`K`$ vs. SNR) The error exponent $`K`$ is monotonically increasing as SNR increases for a given correlation coefficient $`0a<1`$. Moreover, at high SNR the error exponent $`K`$ increases linearly with respect to $`\frac{1}{2}\mathrm{log}[1+\text{SNR}(1a^2)]`$. Proof: See the Appendix. The detrimental effect of correlation at high SNR is clear.Interestingly, the error exponent at high SNR has the same expression as the capacity of the Gaussian channel. The performance degradation due to correlation is equivalent to the SNR decreasing by factor $`(1a^2)`$. The $`\mathrm{log}(1+SNR)`$ increase of $`K`$ w.r.t. SNR is analogous to similar error-rate behavior arising in diversity combining of versions of a communications signal arriving over independent Rayleigh-faded paths in additive noise, where the error probability is given by $`P_e(1+\text{SNR})^L`$ and $`L`$ is the number of independent multipaths. In both cases, the signal component is random. The $`\mathrm{log}`$ SNR behavior of the optimal Neyman-Pearson detector for stochastic signals applies to general correlations as well with a modified definition of SNR. Comparing with the detection of a deterministic signal in noise, where the error exponent is proportional to SNR, the increase of error exponent w.r.t. SNR is much slower for the case of a stochastic signal in noise. Fig. 4 shows the error exponent with respect to SNR for a given correlation strength. The $`\mathrm{log}`$ SNR behavior is evident at high SNR. ## V Extension to The Vector Case In order to treat general cases in which the signal is a higher order AR process or the signal is determined by a linear combination of several underlying phenomena, we now consider a vector state-space model, and extend the results of the previous sections to this model. The hypotheses for the vector case are given by $$\begin{array}{ccc}H_0\hfill & :& y_i=w_i,i=1,2,\mathrm{},n,\hfill \\ H_1\hfill & :& y_i=𝐡^T𝐬_i+w_i,\hfill \end{array}$$ (26) where $`𝐡`$ is a known vector and $`𝐬_i\stackrel{\mathrm{\Delta }}{=}[s_{1i},s_{2i},\mathrm{},s_{mi}]^T`$ is the state of an $`m`$-dimensional process at time $`i`$ following the state-space model $`𝐬_{i+1}`$ $`=`$ $`\mathrm{𝐀𝐬}_i+\mathrm{𝐁𝐮}_i,`$ (27) $`𝐬_1`$ $``$ $`𝒩(\mathrm{𝟎},𝚷_0),`$ $`𝐮_i`$ $`\stackrel{i.i.d.}{}`$ $`𝒩(\mathrm{𝟎},𝐐),𝐐0.`$ We assume that the feedback and input matrices, $`𝐀`$ and $`𝐁`$, are known with $`|\lambda _k(𝐀)|<1`$ for all $`k`$, and the process noise $`\{𝐮_i\}`$ independent of the measurement noise $`\{w_i\}`$. We also assume that the initial state $`𝐬_1`$ is independent of $`𝐮_i`$ for all $`i`$, and the initial covariance $`𝚷_0`$ satisfies the following Lyapunov equation $$𝚷_0=𝐀𝚷_0𝐀^T+\mathrm{𝐁𝐐𝐁}^T.$$ (28) Thus, the signal sequence $`\{𝐬_i\}`$ forms a stationary vector process. In this case the SNR is defined similarly to (5) as $`\frac{𝐡^T𝚷_0𝐡}{\sigma ^2}`$. The autocovariance of the observation sequence $`\{y_i\}`$ is given by $$r_y(i,j)=𝔼\{y_iy_j\}=\{\begin{array}{cc}\sigma ^2\delta _{ij}\hfill & \text{under}H_0,\hfill \\ 𝐡^T𝐀^{|ij|}𝚷_0𝐡+\sigma ^2\delta _{ij}\hfill & \text{under}H_1,\hfill \end{array}$$ (29) where $`\delta _{ij}`$ is the Kronecker delta. Thus, the covariance matrix of the observation under $`H_1`$ is a symmetric Toeplitz matrix with $`𝐡^T𝐀^l𝚷_0𝐡`$ as the $`l`$th off-diagonal entry ($`l1`$). Since $`|\lambda _k(𝐀)|<1`$ for all $`k`$, $`c_l\stackrel{\mathrm{\Delta }}{=}𝐡^T𝐀^l𝚷_0𝐡`$ is an absolutely summable sequence and the eigenvalues of the covariance matrix of $`𝐲_n`$ is bounded both from below and from above. ###### Theorem 4 (Error exponent) For the Neyman-Pearson detector for the hypotheses (26) and (27) with level $`\alpha (0,1)`$ (i.e. $`P_F\alpha `$) and $`|\lambda _k(𝐀)|<1`$ for all $`k`$, the error exponent of the miss probability is given by (21) independently of the value of $`\alpha `$. The steady-state variances of the innovation process $`R_e`$ and $`\stackrel{~}{R}_e`$ calculated under $`H_1`$ and $`H_0`$, respectively, are given by $$R_e=\sigma ^2+𝐡^T\mathrm{𝐏𝐡},$$ (30) where $`𝐏`$ is the unique stabilizing solution of the discrete-time algebraic Riccati equation $$𝐏=\mathrm{𝐀𝐏𝐀}^T+\mathrm{𝐁𝐐𝐁}^T\frac{\mathrm{𝐀𝐏𝐡𝐡}^T\mathrm{𝐏𝐀}^T}{𝐡^T\mathrm{𝐏𝐡}+\sigma ^2},$$ (31) and $$\stackrel{~}{R}_e=\sigma ^2(1+𝐡^T\stackrel{~}{𝐏}𝐡),$$ (32) where $`\stackrel{~}{𝐏}`$ is the unique positive-semidefinite solution of the following Lyapunov equation $$\stackrel{~}{𝐏}=(𝐀𝐊_p𝐡^T)\stackrel{~}{𝐏}(𝐀𝐊_p𝐡^T)^T+𝐊_p𝐊_p^T,$$ (33) and $`𝐊_p=\mathrm{𝐀𝐏𝐡}R_e^1`$. In spectral form, $`K`$ is given by (19), where $`S_y^{(0)}(\omega )=\sigma ^2(\pi \omega \pi )`$ and $`S_y^{(1)}(\omega )`$ is given by $$S_y^{(1)}(\omega )=[𝐡^T(e^{j\omega }𝐈𝐀)^11]\left[\begin{array}{cc}𝐐& 0\\ 0& \sigma ^2\end{array}\right]\left[\begin{array}{c}(e^{j\omega }𝐈𝐀^T)^1𝐡\\ 1\end{array}\right].$$ (34) Proof: See the Appendix. For this vector model, simple results describing the properties of the error exponent are not tractable since the relevant expressions depend on the multiple eigenvalues of the matrix $`𝐀`$. However, (21), (31) and (33) provide closed-form expressions for the error exponent which can easily be explored numerically. ## VI Simulation Results To verify the behavior of the miss probability predicted by our asymptotic analysis, in this section we provide some simulation results. We consider the scalar model (4), for SNR of 10 dB and - 3 dB, and for several correlation strengths. The probability of false alarm is set at 0.1% for all cases we consider. Fig. 5 shows the simulated miss probability as a function of the number of samples for 10 dB SNR. It is seen, as predicted by our analysis, that the i.i.d. case ($`a=0`$) has the largest slope for error decay, and the slope is monotonically decreasing as $`a`$ increases to one. Notice that the error performance for the same number of observations is significantly different for different correlation strengths for the same SNR, and the performance for weak correlation is not much different from the i.i.d. case, as predicted by Fig. 1. It is also seen that the miss probability for the perfectly correlated case ($`a=1`$) is not exponentially decaying, again confirming our analysis. The simulated error performance for SNR of -3 dB is shown in Fig. 6. It is seen that the asymptotic slope of $`\mathrm{log}P_M`$ increases as $`a`$ increases from zero as predicted by Theorem 2, and reaches a maximum with a sudden decrease after the maximum. Notice that the error curve is still not a straight line for the low SNR case due to the $`o(n)`$ term in the exponent of the error probability. Since the error exponent increases only with $`\mathrm{log}`$ SNR, the required number of observations for -3 dB SNR is much larger than for 10 dB SNR for the same miss probability. It is clearly seen that $`P_M`$ is still larger than $`10^2`$ for 200 samples whereas it is $`10^4`$ with 20 samples for the 10 dB SNR case. ## VII Conclusion We have considered the detection of correlated random signals using noisy observations. We have derived the error exponent for the Neyman-Pearson detector of a fixed level using the spectral domain and the innovations approaches. We have also provided the error exponent in closed form for the vector state-space model. The closed-form expression is valid not only for the state-space model but also for any orthogonal transformation of the original observations under the state-space model, since the spectral domain result does not change by orthogonal transformation and Theorem 1 is a closed-form expression of the invariant spectral form. We have investigated the properties of the error exponent for the scalar case. The error exponent is a function of SNR and correlation strength. The behavior of the error exponent as a function of correlation strength is sharply divided into two regimes depending on SNR. For SNR $`1`$ the error exponent is monotonically decreasing in the signal correlation. On the other hand, for SNR $`<1`$, there is a non-zero correlation strength that gives the maximal error exponent. Simulations confirm the validity of our asymptotic results for finite sample sizes. ## Acknowledgement The authors wish to thank an anonymous reviewer for pointing out a simple derivation of the spectral domain result in Section III. ## Appendix Proof of Theorem 1 Since the error exponent for the Neyman-Pearson detector with a fixed level $`\alpha (0,1)`$ is given by the almost-sure limit of the normalized log-likelihood ratio $`\frac{1}{n}\mathrm{log}L_n(𝐲_n)`$ under $`H_0`$ (if the limit exists), we focus on the calculation of the limit. We show that $`\frac{1}{n}\mathrm{log}L_n`$ converges a.s. under $`H_0`$ for Gauss-Markov signals in noise using the limit distribution of the innovations sequence. The log-likelihood ratio is given by $$\mathrm{log}L_n(𝐲_n)=\mathrm{log}p_{1,n}(𝐲_n)\mathrm{log}p_{0,n}(𝐲_n).$$ (35) We have, for the second term on the right-handed side (RHS) of (35), $$p_{0,n}(𝐲_n)=\frac{1}{(2\pi \sigma ^2)^{n/2}}e^{\frac{1}{2}_{i=1}^ny_i^2/\sigma ^2},$$ and so, under $`H_0`$, $`{\displaystyle \frac{1}{n}}\mathrm{log}p_{0,n}(𝐲)`$ $`=`$ $`{\displaystyle \frac{1}{2}}\mathrm{log}(2\pi \sigma ^2){\displaystyle \frac{1}{2\sigma ^2n}}{\displaystyle \underset{i=1}{\overset{n}{}}}y_i^2,`$ (36) $``$ $`{\displaystyle \frac{1}{2}}\mathrm{log}(2\pi \sigma ^2){\displaystyle \frac{1}{2}}\text{a.s.},`$ (37) since $`\frac{1}{n}_{i=1}^ny_i^2𝔼_0\{Y_1^2\}=\sigma ^2`$ a.s. under $`H_0`$ as $`n\mathrm{}`$ by the SLLN. Now consider the first term on the RHS of (35). The log-likelihood under $`H_1`$ can be obtained via the Kalman recursion for the innovations ,. Specifically, define $`l_i\stackrel{\mathrm{\Delta }}{=}\mathrm{log}p_{1,i}(y_1,y_2,\mathrm{},y_i)`$ and $`y_1^i\stackrel{\mathrm{\Delta }}{=}\{y_1,y_2,\mathrm{},y_i\}`$. Then, $$p_{1,i}(y_1^i)=p_{1,i}(y_1^{i1})p_{1,i}(y_i|y_1^{i1}),2in.$$ (38) Hence, $$l_i=l_{i1}+\mathrm{log}p_{1,i}(y_i|y_1^{i1}),2in,$$ (39) where $`l_1=\mathrm{log}p_{1,1}(y_1)`$. Since the joint distribution of $`\{y_1,y_2,\mathrm{},y_i\}`$ is Gaussian, the conditional distribution $`p_{1,i}(y_i|y_1^{i1})`$ is also Gaussian with mean $`\widehat{y}_{i|i1}`$ and variance $`R_{e,i}`$. $`l_i`$ is expressed using the innovations representation by $$l_i=l_{i1}\frac{1}{2}\mathrm{log}(2\pi R_{e,i})\frac{1}{2}\frac{e_i^2}{R_{e,i}},$$ (40) where the minimum mean square error (MMSE) prediction $`\widehat{y}_{i|i1}`$ of $`y_i`$ is the conditional expectation $`𝔼_1\{y_i|y_1^{i1}\}`$ and the innovation is given by $`e_i\stackrel{\mathrm{\Delta }}{=}y_i\widehat{y}_{i|i1}`$ with variance $`R_{e,i}=𝔼_1\{e_i^2\}`$. Hence, $$\frac{1}{n}\mathrm{log}p_{1,n}(𝐲_n)=\frac{l_n}{n}=\frac{1}{2}\mathrm{log}(2\pi )\frac{1}{2n}\underset{i=1}{\overset{n}{}}\mathrm{log}R_{e,i}\frac{1}{2n}\underset{i=1}{\overset{n}{}}\frac{e_i^2}{R_{e,i}}.$$ (41) The second term on the RHS of (41) is not random, and we have $$\frac{1}{2n}\underset{i=1}{\overset{n}{}}\mathrm{log}R_{e,i}\frac{1}{2}\mathrm{log}R_e,$$ (42) by the Cesáro mean theorem since $`R_{e,i}R_e`$ and $`R_{e,i}\sigma ^2>0`$ for all $`i`$ where $`R_e`$ is given by $$R_e=P+\sigma ^2,$$ (43) and where $`P`$ is the steady-state error variance of the optimal one-step predictor for the signal $`\{s_i\}`$. Now representing $`e_i`$ as a linear combination of $`y_1,y_2,\mathrm{},y_i`$ gives $`e_i`$ $`=`$ $`y_iK_{p,i1}y_{i1}(aK_{p,i1})K_{p,i2}y_{i2}(aK_{p,i1})(aK_{p,i2})K_{p,i3}y_{i3}`$ (44) $`(aK_{p,i1})(aK_{p,i2})(aK_{p,i3})K_{p,i4}y_{i4}\mathrm{},`$ where $`K_{p,i}\stackrel{\mathrm{\Delta }}{=}aP_iR_{e,i}^1`$ is the Kalman prediction gain, $`P_i\stackrel{\mathrm{\Delta }}{=}𝔼_1\{(s_i\widehat{s}_{i|i1})^2\}`$ is the error variance at time $`i`$, $`\widehat{s}_{i|i1}`$ is the linear MMSE prediction of $`s_i`$ given $`y_1^{i1}`$. Since the Kalman filter converges asymptotically to the time-invariant recursive Wiener filter for $`0a<1`$, we have asymptotically $`e_i`$ $`=`$ $`y_iK_py_{i1}(aK_p)K_py_{i2}(aK_p)(aK_p)K_py_{i3}`$ (45) $`(aK_p)(aK_p)(aK_p)K_py_{i4}\mathrm{},`$ where $`K_p`$ is the steady-state Kalman prediction gain. Thus, under $`H_0`$ the innovations sequence becomes the output of a stable recursive filter driven by an i.i.d. sequence $`\{y_i\}`$, and it is known to be an ergodic sequence. By the ergodic theorem, $`\frac{1}{n}_{i=1}^ne_i^2`$ converges to the true expectation, which is given by $$\stackrel{~}{R}_e=\underset{i\mathrm{}}{lim}𝔼_0\{e_i^2\}=\sigma ^2\left(1+\frac{K_p^2}{1(aK_p)^2}\right)$$ (46) since $`\{y_i\}_{i=1}^{\mathrm{}}`$ is an independent sequence under $`H_0`$. Substituting $`K_p`$ and $`R_e`$, we have $$\stackrel{~}{R}_e=\sigma ^2\left(1+\frac{a^2P^2}{R_e^2a^2\sigma ^4}\right)=\sigma ^2\left(1+\frac{a^2P^2}{P^2+2\sigma ^2P+(1a^2)\sigma ^4}\right).$$ (47) Now, the last term on the RHS of (41) is given by $`{\displaystyle \frac{1}{2n}}{\displaystyle \underset{i=1}{\overset{n}{}}}{\displaystyle \frac{e_i^2}{R_{e,i}}}`$ $`=`$ $`{\displaystyle \frac{1}{2n}}{\displaystyle \underset{i=1}{\overset{n}{}}}{\displaystyle \frac{e_i^2}{R_e}}{\displaystyle \frac{R_e}{R_{e,i}}}={\displaystyle \frac{1}{2n}}{\displaystyle \underset{i=1}{\overset{n}{}}}{\displaystyle \frac{e_i^2}{R_e}}{\displaystyle \frac{R_e+Cϵ^nCϵ^n}{R_e+Cϵ^n}},`$ (48) $`=`$ $`{\displaystyle \frac{1}{2nR_e}}{\displaystyle \underset{i=1}{\overset{n}{}}}e_i^2{\displaystyle \frac{1}{2nR_e}}{\displaystyle \underset{i=1}{\overset{n}{}}}e_i^2{\displaystyle \frac{Cϵ^n}{R_e+Cϵ^n}},`$ (49) where some positive constant $`C`$ and $`|ϵ|<1`$ by the exponential convergence of $`R_{e,i}`$ to $`R_e`$. The first term in (49) converges to $`\frac{\stackrel{~}{R}_e}{2R_e}`$ and the second term converges to zero since $`_{i=1}^ne_i^2/n`$ converges to a finite constant and $`R_e\sigma ^2>0`$. Hence, (21) follows for $`0a<1`$. When $`a=1`$, we have $`P_M\mathrm{\Theta }\left(\frac{1}{\sqrt{n}}\right)`$ and $`K=0`$. We also have $`P=0`$, $`\stackrel{~}{R}_e=R_e=\sigma ^2`$ in (22) - (24) at $`a=1`$. Thus, (20) has a value of zero at $`a=1`$, and Theorem 1 holds for $`0a1`$. Now we show that (21) is equivalent to the spectral domain result (19) using spectral factorization. From the spectral domain form (19) we have $`K`$ $`=`$ $`{\displaystyle \frac{1}{4\pi }}{\displaystyle _0^{2\pi }}\mathrm{log}{\displaystyle \frac{S_y^{(1)}(\omega )}{\sigma ^2}}d\omega +{\displaystyle \frac{1}{4\pi }}{\displaystyle _0^{2\pi }}{\displaystyle \frac{\sigma ^2}{S_y^{(1)}(\omega )}}𝑑\omega {\displaystyle \frac{1}{2}},`$ (50) $`=`$ $`{\displaystyle \frac{1}{4\pi }}{\displaystyle _0^{2\pi }}\mathrm{log}S_y^{(1)}(\omega )𝑑\omega +{\displaystyle \frac{1}{4\pi }}{\displaystyle _0^{2\pi }}{\displaystyle \frac{\sigma ^2}{S_y^{(1)}(\omega )}}𝑑\omega {\displaystyle \frac{1}{2}}\mathrm{log}\sigma ^2{\displaystyle \frac{1}{2}}.`$ First, consider the first term on the RHS of (50). The argument of the logarithm is the power spectral density of the observation sequence $`\{y_i\}`$ under $`H_1`$. From Wiener filtering theory, the canonical spectral factorization for $`S_y^{(1)}(z)`$ is given by (, p.275) $`S_y^{(1)}(z)`$ $`=`$ $`L(z)R_eL^{}(z^{}),`$ (51) where $`L^1(z)`$ is the whitening filter. Hence, we have $`{\displaystyle \frac{1}{4\pi }}{\displaystyle _0^{2\pi }}\mathrm{log}S_y^{(1)}(\omega )𝑑\omega `$ $`=`$ $`{\displaystyle \frac{1}{4\pi }}{\displaystyle _0^{2\pi }}\mathrm{log}(L(e^{j\omega })R_eL^{}(e^{j\omega }))𝑑\omega ,`$ $`=`$ $`{\displaystyle \frac{1}{4\pi }}{\displaystyle _0^{2\pi }}(\mathrm{log}R_e+\mathrm{log}L(e^{j\omega })+\mathrm{log}L^{}(e^{j\omega }))𝑑\omega ,`$ $`=`$ $`{\displaystyle \frac{1}{2}}\mathrm{log}R_e,`$ where the last step follows from the cancellation of two other terms in para-Hermitian conjugacy. Now, consider the second term on the RHS of (50). From (51), we have $$\frac{\sigma ^2}{S_y^{(1)}(z)}=\frac{\sigma ^2L^1(z)(L^{}(z^{}))^1}{R_e},$$ (52) which is the spectral density of the innovations process under $`H_0`$ divided by $`R_e`$, since $`\{y_i\}`$ is an i.i.d. sequence with variance $`\sigma ^2`$ under $`H_0`$ and $`L^1(z)`$ is the whitening filter. Since the variance of a stationary process is given by the autocovariance function $`r(l)`$ setting $`l=0`$, we have, by the definition of $`\stackrel{~}{R}_e`$, $$\stackrel{~}{R}_e=r(0)=\frac{1}{2\pi }_0^{2\pi }\sigma ^2[L^1(z)(L^{}(z^{}))^1]_{z=e^{j\omega }}𝑑\omega $$ (53) since the spectral density is the Fourier transform of the autocovariance function. (Eq. (23) is an explicit formula for (53).) Hence, we have $$\frac{1}{4\pi }_0^{2\pi }\frac{\sigma ^2}{S_y^{(1)}(\omega )}𝑑\omega =\frac{1}{2}\frac{\stackrel{~}{R}_e}{R_e},$$ (54) and (50) is given by $$\frac{1}{2}\mathrm{log}R_e+\frac{1}{2}\frac{\stackrel{~}{R}_e}{R_e}\frac{1}{2}\mathrm{log}\sigma ^2\frac{1}{2}=\frac{1}{2}\mathrm{log}\frac{R_e}{\sigma ^2}+\frac{1}{2}\frac{\stackrel{~}{R}_e}{R_e}\frac{1}{2},$$ (55) which is the error exponent in Theorem 1 that we derived using the innovations approach. $`\mathrm{}`$ ###### Lemma 1 The partial derivative of the error exponent with respect to the correlation coefficient $`a`$ is given by $$\frac{K}{a}=\frac{\mathrm{\Gamma }(ba)}{r_e(1b^2)}\left(\frac{1}{1ab}\frac{2(1ab)}{r_e(1b^2)^2}\right)$$ (56) for a fixed SNR $`\mathrm{\Gamma }`$, where $`b=a/r_e`$ and $`r_e=R_e/\sigma ^2`$. It is easily seen that the partial derivative $`\frac{K}{a}`$ is a continuous function of $`a`$ for $`0a<1`$ since $`r_e`$ is a continuous function of $`a`$ from (22) and (24). Proof of Lemma 1 We use the spectral domain form for the error exponent. $`K`$ $`=`$ $`{\displaystyle \frac{1}{4\pi }}{\displaystyle _0^{2\pi }}\mathrm{log}{\displaystyle \frac{\sigma ^2}{\sigma ^2+S_s(\omega )}}d\omega +{\displaystyle \frac{1}{4\pi }}{\displaystyle _0^{2\pi }}{\displaystyle \frac{\sigma ^2}{\sigma ^2+S_s(\omega )}}𝑑\omega {\displaystyle \frac{1}{2}},`$ (57) $`=`$ $`{\displaystyle \frac{1}{4\pi }}{\displaystyle _0^{2\pi }}\mathrm{log}{\displaystyle \frac{1}{1+\mathrm{\Gamma }\stackrel{~}{S}_s(\omega )}}d\omega +{\displaystyle \frac{1}{4\pi }}{\displaystyle _0^{2\pi }}{\displaystyle \frac{1}{1+\mathrm{\Gamma }\stackrel{~}{S}_s(\omega )}}𝑑\omega {\displaystyle \frac{1}{2}},`$ where $$S_s(\omega )=\frac{\mathrm{\Pi }_0(1a^2)}{12a\mathrm{cos}\omega +a^2},\stackrel{~}{S}_s(\omega )=S_s(\omega )/\mathrm{\Pi }_0,\mathrm{\Gamma }=\frac{\mathrm{\Pi }_0}{\sigma ^2}.$$ (58) The spectral density of the observation sequence $`\{y_i\}`$ is given by $`S_y^{(1)}(z)`$ $`=`$ $`\sigma ^2+S_s(z)=\sigma ^2\left(1+{\displaystyle \frac{Q/\sigma ^2}{(1az^1)(1az)}}\right),`$ (59) where $`Q=\mathrm{\Pi }_0(1a^2)`$, and its canonical spectral factorization is given by (, p.242) $$S_y^{(1)}(z)=\sigma ^2L(z)r_eL^{}(z^{})=\sigma ^2r_e\frac{1bz}{1az}\frac{1bz^1}{1az^1},$$ (60) where $`b=a/r_e`$ ($`|a|<1`$ and $`|b|<1`$) and $$r_e=\frac{\sqrt{[1+a^2+Q/\sigma ^2]^24a^2}+1+a^2+Q/\sigma ^2}{2}=\frac{R_e}{\sigma ^2}.$$ (61) The partial derivative of $`K`$ with respect to $`a`$ is given by $$\frac{K}{a}=\frac{1}{4\pi }_0^{2\pi }\frac{\mathrm{\Gamma }\stackrel{~}{S}_s^{}(\omega )}{1+\mathrm{\Gamma }\stackrel{~}{S}_s(\omega )}𝑑\omega \frac{1}{4\pi }_0^{2\pi }\frac{\mathrm{\Gamma }\stackrel{~}{S}_s^{}(\omega )}{(1+\mathrm{\Gamma }\stackrel{~}{S}_s(\omega ))^2}𝑑\omega ,$$ (62) where $$\stackrel{~}{S}_s^{}(\omega )=\frac{\stackrel{~}{S}_s(\omega )}{a}=\frac{2[(1+a^2)\mathrm{cos}\omega 2a]}{(12a\mathrm{cos}\omega +a^2)^2}.$$ (63) Consider the first term on the RHS of (62). Using the canonical spectral decomposition (60), we have $`{\displaystyle \frac{1}{4\pi }}{\displaystyle _0^{2\pi }}{\displaystyle \frac{\mathrm{\Gamma }\stackrel{~}{S}_s^{}(\omega )}{1+\mathrm{\Gamma }\stackrel{~}{S}_s(\omega )}}𝑑\omega `$ $`=`$ $`{\displaystyle \frac{1}{4\pi }}{\displaystyle _0^{2\pi }}{\displaystyle \frac{\frac{2\mathrm{\Gamma }[(1+a^2)\mathrm{cos}\omega 2a]}{(12a\mathrm{cos}\omega +a^2)^2}}{r_e\frac{1be^{j\omega }}{1ae^{j\omega }}\frac{1be^{j\omega }}{1ae^{j\omega }}}}𝑑\omega ,`$ (64) $`=`$ $`{\displaystyle \frac{1}{4\pi }}{\displaystyle \frac{\mathrm{\Gamma }[(1+a^2)(z+z^1)4a]}{r_e(1az)(1az^1)(1bz)(1bz^1)}\frac{dz}{jz}},`$ $`=`$ $`{\displaystyle \frac{1}{4\pi j}}{\displaystyle \frac{\mathrm{\Gamma }[(1+a^2)(z^2+1)4az]}{r_e(1az)(za)(1bz)(zb)}𝑑z},`$ $`=`$ $`{\displaystyle \frac{1}{4\pi j}}{\displaystyle \frac{\mathrm{\Gamma }}{r_e}}2\pi j{\displaystyle \underset{|z|<1}{}}\text{Residues of integrand},`$ $`=`$ $`{\displaystyle \frac{\mathrm{\Gamma }}{2r_e}}\left({\displaystyle \frac{1a^2}{(1ab)(ab)}}+{\displaystyle \frac{(1+a^2)(1+b^2)4ab}{(1ab)(ba)(1b^2)}}\right),`$ $`=`$ $`{\displaystyle \frac{\mathrm{\Gamma }}{r_e}}{\displaystyle \frac{(ba)}{(1ab)(1b^2)}},`$ where we have substituted $`z=e^{j\omega }`$, and used the residue theorem. The second term on the RHS of (62) is similarly obtained: $`{\displaystyle \frac{1}{4\pi }}{\displaystyle _0^{2\pi }}{\displaystyle \frac{\mathrm{\Gamma }\stackrel{~}{S}_s^{}(\omega )}{(1+\mathrm{\Gamma }\stackrel{~}{S}_s(\omega ))^2}}𝑑\omega `$ $`=`$ $`{\displaystyle \frac{1}{4\pi }}{\displaystyle \frac{\mathrm{\Gamma }[(1+a^2)(z+z^1)4a]}{r_e^2(1bz)^2(1bz^1)^2}\frac{dz}{jz}},`$ $`=`$ $`{\displaystyle \frac{\mathrm{\Gamma }}{4\pi r_e^2j}}{\displaystyle \frac{(1+a^2)(z^2+1)4az}{(1bz)^2(zb)^2}𝑑z},`$ $`=`$ $`{\displaystyle \frac{\mathrm{\Gamma }}{4\pi r_e^2j}}2\pi j\text{Res}(z=b),`$ where $$\text{Res}(z=b)=\left[\frac{}{z}\frac{(1+a^2)(z^2+1)4az}{(1bz)^2}\right]_{z=b}=\frac{4(ba)(1ab)}{(1b^2)^3}.$$ (65) Hence, $`K^{}`$ is given by $$\frac{K}{a}=\frac{\mathrm{\Gamma }(ba)}{r_e(1b^2)}\left(\frac{1}{1ab}\frac{2(1ab)}{r_e(1b^2)^2}\right).$$ (66) $`\mathrm{}`$ Proof of Theorem 2 First, the continuity of $`K`$ in (21) is straightforward as a function of $`R_e`$ and $`\stackrel{~}{R}_e`$ since $`R_e\sigma ^2`$, and the continuity of $`R_e`$ and $`\stackrel{~}{R}_e`$ is also trivial as a function of $`a`$ and $`P`$ from (22) and (23). Thus, we need only to show the continuity of $`P`$ as a function of $`a`$, i.e., the nonnegativity of the argument of the square root in (24). The argument can be rewritten as $$[\sigma ^2(1a^2)Q]^2+4\sigma ^2Q=\sigma ^4(1a^2)[(1a^2)(1\mathrm{\Gamma })^2+4\mathrm{\Gamma }],$$ (67) which is nonnegative if either $`\mathrm{\Gamma }=1`$ or $`\frac{4\mathrm{\Gamma }}{(1\mathrm{\Gamma })^2}\mathrm{min}_{a[0,1]}(1a^2)=0`$ if $`\mathrm{\Gamma }1`$. Thus, $`K`$ is a continuous function of $`a`$ $`(0a1)`$ for any SNR $`\mathrm{\Gamma }0`$. (i) SNR $`1`$: Since $`b=a/r_e`$ in (66), we have $`{\displaystyle \frac{K}{a}}`$ $`=`$ $`{\displaystyle \frac{\mathrm{\Gamma }(ba)}{r_e(1b^2)}}\left({\displaystyle \frac{1}{1ab}}{\displaystyle \frac{2b(1ab)}{a(1b^2)^2}}\right),`$ (68) $`=`$ $`{\displaystyle \frac{\mathrm{\Gamma }(ba)}{r_e(1b^2)}}{\displaystyle \frac{[a(1b^2)^22b(1ab)^2)]}{a(1ab)(1b^2)^2}},`$ $`=`$ $`{\displaystyle \frac{\mathrm{\Gamma }(ba)}{r_e(1b^2)}}{\displaystyle \frac{[a(1+b^2)^22b(1+a^2b^2)]}{a(1ab)(1b^2)^2}},`$ $`=`$ $`{\displaystyle \frac{\mathrm{\Gamma }(ba)[br_e(1+b^2)^22b(1+a^2b^2)]}{r_ea(1ab)(1b^2)^3}},`$ $`=`$ $`{\displaystyle \frac{\mathrm{\Gamma }(ba)br_e^1[(1+\stackrel{~}{Q}+a^2)^22r_e(1+a^4r_e^2)]}{r_ea(1ab)(1b^2)^3}},`$ where $`\stackrel{~}{Q}\stackrel{\mathrm{\Delta }}{=}\mathrm{\Gamma }(1a^2)`$ and we have used the relation $`r_e(1+b^2)=1+\stackrel{~}{Q}+a^2`$ in the canonical spectral factorization. (See p.242.) We also have the relation $$r_e+\frac{a^2}{r_e}=1+\stackrel{~}{Q}+a^2,$$ (69) which implies $$r_e(1+a^4r_e^2)1+\stackrel{~}{Q}+a^2,$$ (70) since $`0a<1`$. Hence, for the last term in the numerator of (68) we have $$(1+\stackrel{~}{Q}+a^2)^22r_e(1+a^4r_e^2)(1+\stackrel{~}{Q}+a^2)^22(1+\stackrel{~}{Q}+a^2).$$ (71) The RHS of (71) is positive for $`1+\stackrel{~}{Q}+a^2=1+a^2+\mathrm{\Gamma }(1a^2)2`$, which reduces to the condition $`\mathrm{\Gamma }1`$. Since $$r_e=R_e/\sigma ^2=1+P/\sigma ^2>1\text{for}0a<1,$$ (72) we have $$ba=(r_e^11)a<0.$$ (73) Hence, $`\frac{K}{a}0`$ for $`0a<1`$ and $`\mathrm{\Gamma }1`$, and $`K`$ is monotonically decreasing as $`a1`$ for $`\mathrm{\Gamma }1`$. (ii) SNR $`<1`$: For a given $`\mathrm{\Gamma }`$, denote the last term in the numerator of (68) by $$f_\mathrm{\Gamma }(a)\stackrel{\mathrm{\Delta }}{=}(1+\stackrel{~}{Q}+a^2)^22r_e(1+a^4r_e^2).$$ (74) Then, we can write $`f_\mathrm{\Gamma }(a=0)`$ $`=`$ $`(1+\mathrm{\Gamma })^22(1+\mathrm{\Gamma })=\mathrm{\Gamma }^21,`$ (75) since $`\stackrel{~}{Q}=\mathrm{\Gamma }(1a^2)`$ and $`r_e=1+\mathrm{\Gamma }`$ for $`a=0`$ from (61). We have $`f_\mathrm{\Gamma }(a=0)<0`$ for $`\mathrm{\Gamma }<1`$ and $`\frac{K}{a}>0`$ from $`(\text{68})`$ since $`ba<0`$. Hence, $`K`$ increases as $`a`$ increases in the neighborhood of $`a=0`$ with $`K(a=0)=D(𝒩(0,1)||𝒩(0,1+\mathrm{\Gamma })>0`$ if $`0<\mathrm{\Gamma }<1`$. However, $`K0`$ as $`a`$ approaches one since $`P_M\mathrm{\Theta }(\frac{1}{\sqrt{n}})`$ at $`a=1`$. Hence, $`K`$ achieves a maximum at nonzero $`a`$ for SNR $`\mathrm{\Gamma }<1`$ since $`K`$ is a continuous function of $`a`$, and the value of $`a`$ achieving the maximum is given by $`f_\mathrm{\Gamma }(a)=0`$ since $`\frac{K}{a}`$ is also continuous with $`a`$. As SNR $`\mathrm{\Gamma }0`$, we have $$\stackrel{~}{Q}=\mathrm{\Gamma }(1a^2)0\text{and}r_e=1+P/\sigma ^21.$$ (76) The last term in the numerator of (68) is given by $`(1+\stackrel{~}{Q}+a^2)^22r_e(1+a^4r_e^2)`$ $``$ $`(1+a^2)^22(1+a^4),`$ (77) $`=`$ $`(1a^2)^2<0,`$ for $`0a<1`$ as $`\mathrm{\Gamma }0`$. Hence, for any $`\delta >0`$, there exists $`\mathrm{\Gamma }_0`$ small enough such that for all $`\mathrm{\Gamma }<\mathrm{\Gamma }_0(\delta )`$, $`(1+\stackrel{~}{Q}+a^2)^22r_e(1+a^4r_e^2)(1a^2)^2+\delta `$. This guarantees that for $`\mathrm{\Gamma }<\mathrm{\Gamma }_0(\delta ),K/a>0`$ for $`0<a<\sqrt{1\sqrt{\delta }}`$, and $`a^{}\sqrt{1\sqrt{\delta }}`$. $`\mathrm{}`$ Proof of Theorem 3 Let $`s=1+\mathrm{\Gamma }\stackrel{~}{S}_s(\omega )`$ where $`\stackrel{~}{S}_s(\omega )`$ is given by (58). Then, from (57), the partial derivative of $`K`$ w.r.t. $`\mathrm{\Gamma }`$ is given by $$\frac{K}{\mathrm{\Gamma }}=\frac{1}{2\pi }_0^{2\pi }\frac{}{s}\left(\frac{1}{2}\mathrm{log}\frac{1}{s}+\frac{1}{2s}\frac{1}{2}\right)\frac{s}{\mathrm{\Gamma }}𝑑\omega ,$$ (78) where $$\frac{}{s}\left(\frac{1}{2}\mathrm{log}\frac{1}{s}+\frac{1}{2}\frac{1}{s}\frac{1}{2}\right)=\frac{1}{2}\frac{s1}{s^2}=\frac{1}{2}\frac{\mathrm{\Gamma }\stackrel{~}{S}_s(\omega )}{s^2}>0,$$ (79) and $$\frac{s}{\mathrm{\Gamma }}=\stackrel{~}{S}_s(\omega )=\frac{1a^2}{12a\mathrm{cos}(\omega )+a^2}>0$$ (80) for $`0a<1`$. Hence, $$\frac{K}{\mathrm{\Gamma }}>0,$$ (81) and the error exponent $`K`$ increases monotonically as SNR increases for a given $`a`$ ($`0a<1`$). At high SNR, we have $`P`$ $``$ $`Q,`$ $`R_e`$ $``$ $`Q+\sigma ^2,`$ $`\stackrel{~}{R}_e`$ $``$ $`\sigma ^2(1+a^2).`$ Hence, from (21), the error exponent is given at high SNR by $`K`$ $``$ $`{\displaystyle \frac{1}{2}}\mathrm{log}{\displaystyle \frac{Q+\sigma ^2}{\sigma ^2}}+{\displaystyle \frac{1}{2}}{\displaystyle \frac{\sigma ^2(1+a^2)}{Q+\sigma ^2}},`$ (82) $`=`$ $`{\displaystyle \frac{1}{2}}\mathrm{log}{\displaystyle \frac{Q+\sigma ^2}{\sigma ^2}}+{\displaystyle \frac{1}{2}}{\displaystyle \frac{1+a^2}{\frac{Q+\sigma ^2}{\sigma ^2}}},Q=\mathrm{\Pi }_0(1a^2),`$ $`=`$ $`{\displaystyle \frac{1}{2}}\mathrm{log}\left(1+{\displaystyle \frac{\mathrm{\Pi }_0}{\sigma ^2}}(1a^2)\right)+{\displaystyle \frac{1}{2}}{\displaystyle \frac{1+a^2}{1+\frac{\mathrm{\Pi }_0}{\sigma ^2}(1a^2)}}.`$ Since the first term is dominant at high SNR, the theorem follows. $`\mathrm{}`$ Proof of Theorem 4 Since the error exponent is given by the asymptotic Kullback-Leibler rate (3) and its representation by innovations for the vector case is the same as (36) and (41). We need only to calculate $`R_e`$ and $`\stackrel{~}{R}_e`$ for the vector model. The steady-state variance $`R_e`$ for the innovations under $`H_1`$ is given by the conventional result of the state-space model $$R_e=𝐡^T\mathrm{𝐏𝐡}+\sigma ^2,$$ (83) and $`𝐏`$ is the unique Hermitian solution of the discrete-time Riccati equation $$𝐏=\mathrm{𝐀𝐏𝐀}^T+\mathrm{𝐁𝐐𝐁}^T\frac{\mathrm{𝐀𝐏𝐡𝐡}^T\mathrm{𝐏𝐀}^T}{𝐡^T\mathrm{𝐏𝐡}+\sigma ^2},$$ (84) such that $`𝐀𝐊_p𝐡^T`$ is stable (the existence of the solution is guaranteed since $`𝐀`$ is stable, $`\text{diag}(𝐐,\sigma ^2)0`$, and $`S_y^{(1)}(\omega )>0`$ due to the additive noise. See p. 277.), where $$𝐊_p=\mathrm{𝐀𝐏𝐡}R_e^1.$$ (85) For $`\stackrel{~}{R}_e`$, we again represent $`e_i`$ as a linear combination of $`y_1,y_2,\mathrm{},y_i`$, and $`e_i`$ is given by $`e_i`$ $`=`$ $`y_i𝐡^T𝐊_{p,i1}y_{i1}𝐡^T(𝐀𝐊_{p,i1}𝐡^T)𝐊_{p,i2}y_{i2}`$ (86) $`𝐡^T(𝐀𝐊_{p,i1}𝐡^T)(𝐀𝐊_{p,i2}𝐡^T)𝐊_{p,i3}y_{i3}`$ $`𝐡^T(𝐀𝐊_{p,i1}𝐡^T)(𝐀𝐊_{p,i2}𝐡^T)(𝐀𝐊_{p,i3}𝐡^T)𝐊_{p,i4}y_{i4}\mathrm{}`$ where $`𝐊_{p,i}`$ is the Kalman prediction gain given by $`𝐊_{p,i}=\mathrm{𝐀𝐏}_i𝐡/(𝐡^T𝐏_i𝐡+\sigma ^2)`$ with the one-step prediction error covariance matrix $`𝐏_i`$. Since the Kalman filter converges asymptotically to the time-invariant recursive Wiener filter when $`𝐀`$ is stable, we have asymptotically $`e_i`$ $`=`$ $`y_i𝐡^T𝐊_py_{i1}𝐡^T(𝐀𝐊_p𝐡^T)𝐊_py_{i2}`$ (87) $`𝐡^T(𝐀𝐊_p𝐡^T)(𝐀𝐊_p𝐡^T)𝐊_py_{i3}`$ $`𝐡^T(𝐀𝐊_p𝐡^T)(𝐀𝐊_p𝐡^T)(𝐀𝐊_p𝐡^T)𝐊_py_{i4}\mathrm{}`$ where $`𝐊_p`$ is the steady-state Kalman prediction gain. Thus, the innovation sequence becomes the output of a stable recursive filter driven by the i.i.d. sequence $`\{y_i\}`$ under $`H_0`$ as in the scalar case, and the ergodic theorem holds for $`\frac{1}{n}_{i=1}^ne_i^2`$. $`\stackrel{~}{R}_e`$ $`=`$ $`\underset{i\mathrm{}}{lim}𝔼_0\{e_i^2\},`$ (88) $`=`$ $`\sigma ^2+\sigma ^2𝐡^T\left({\displaystyle \underset{k=0}{\overset{\mathrm{}}{}}}(𝐀𝐊_p𝐡^T)^k𝐊_p𝐊_p^T[(𝐀𝐊_p𝐡^T)^k]^T\right)𝐡.`$ Let $`\stackrel{~}{𝐏}`$ be defined as $$\stackrel{~}{𝐏}\stackrel{\mathrm{\Delta }}{=}\underset{k=0}{\overset{\mathrm{}}{}}(𝐀𝐊_p𝐡^T)^k𝐊_p𝐊_p^T[(𝐀𝐊_p𝐡^T)^k]^T.$$ (89) $`\stackrel{~}{𝐏}`$ is finite since $`𝐀𝐊_p𝐡^T`$ is stable by the property of the solution of (84), and is given by the unique solution of the following Lyapunov equation $$\stackrel{~}{𝐏}(𝐀𝐊_p𝐡^T)\stackrel{~}{𝐏}(𝐀𝐊_p𝐡^T)^T=𝐊_p𝐊_p^T.$$ (90) (Since $`𝐀𝐊_p𝐡^T`$ is stable and $`𝐊_p𝐊_p^T0`$, there exists a unique, Hermitian, and positive semi-definite solution $`\stackrel{~}{𝐏}`$ of (90).) The spectrum for the vector model is given by (34) . $`\mathrm{}`$
warning/0506/nlin0506020.html
ar5iv
text
# FOR SUBMISSION TO PHYSICA D. Today’s version Clumps and patches in self-aggregation of finite size particles ## 1 Historical perspective of continuum models of self-aggregation Many fields of physics, chemistry and biology deal with the problem of describing the evolution of the macroscopic density of a large number of particles, which self-consistently attract each other over distances large compared to their mean separation. A related paradigm arises in biosciences, particularly in chemotaxis: the study of the influence of chemical substances in the environment on the motion of mobile species which secrete these substances. One of the most famous among such models is the Keller-Segel elliptic-parabolic system of partial differential equations , which was introduced to explore the effects of nonlinear cross diffusion in the formation of aggregates and patterns by chemotaxis in the aggregation of the slime mold Dictyostelium discoidium. The Keller-Segel model consists of two strongly coupled reaction-diffusion equations, $$\rho _t+\text{div}𝐉=0𝐉=\rho \mu (\mathrm{\Phi })\mathrm{\Phi }D\rho ϵ\mathrm{\Phi }_t+L\mathrm{\Phi }=\gamma \rho ,$$ (1) expressing the coupled evolution of concentration of organisms (density) $`\rho `$ and concentration of chemotactic agent (potential) $`\mathrm{\Phi }`$. These challenging nonlinear equations were posed for chemotaxis as an initial value problem with Neumann boundary conditions initial data. The constants $`\gamma ,D,ϵ>0`$ are assumed to be positive and $`𝐉`$ is the flux of organisms (particles). The linear operator $`L`$ is taken to be positive and symmetric. For example, one may choose it to be the Laplacian $`L=\mathrm{\Delta }`$ or the Helmholtz operator, $`L=1\alpha ^2\mathrm{\Delta }`$. For the moment, the functional dependence of the mobility $`\mu `$ will be left unspecified.<sup>1</sup><sup>1</sup>1One recovers the precise expression of Keller and Segel by rewriting the mobility term as $`\mu (\mathrm{\Phi })\mathrm{\Phi }=\chi (\mathrm{\Phi })`$, where $`\chi (\mathrm{\Phi })`$ is the “sensitivity function.” The development below introduces a different functional dependence in the mobility $`\mu `$. The “simplified” Keller-Segel model is obtained by setting $`ϵ=0`$ in (1). Because of its fascinating mathematical properties – such as finite-time concentration of the density $`\rho `$ into Dirac delta-function singularities, starting from smooth initial conditions – and its relevance to chemotaxis, the Keller-Segel system has attracted sustained interest in mathematical biology. To avoid excessive citation here, we shall only refer to the review article where over 150 references on the role of (1) in the chemotaxis problem may be found. The applications of equations in the form of system (1) in physics and chemistry may be traced back into the 19th century. The history of these equations is a story of recurrent rediscovery, each time as a leading order description of singularity formation, occurring as the tendency to diffuse is eventually overwhelmed by long-range forces of attraction. Quite likely, their story has not finished yet and these equations will be needed again for the insights they provide into nonlinear multiscale dynamics in the competition between coalescence and diffusion. Three main modeling steps figure in the derivation of (1) and similar equations. First, the flux $`𝐉`$ must be modeled in terms of the local particle density, its spatial gradient and the gradient of the potential. Second, an equation must be formulated for determining the potential $`\mathrm{\Phi }`$ from the local density $`\rho `$. Third, conservation of mass, charge, or particle number must be introduced as the evolution equation for density. Historically, it seems that Debye and Hückel in 1923 were the first to put all three of these modeling steps together. They derived the evolutionary system (1) in their article on the theory of electrolytes. In particular, the simplified model with $`ϵ=0`$ in (1) may be found as equations (2) and (2$`^{}`$) in Debye and Hückel . Consequently, the simplified evolutionary system (1) with $`ϵ=0`$ may also be called the Debye-Hückel equations. The Debye-Hückel equations (1) with $`ϵ=0`$ also appeared very early in astrophysics as the Poisson-Smoluchowski system, a particular Fokker-Planck equation describing a self-attracting reinforced random walk of particles (with mass, but without inertia) diffusing by local collisions and drifting against friction under their mutual long-range attraction. As is well known, gravitational force leads to collapse in this case, when mutual attraction finally prevails over diffusion and friction. See Chandrasekhar (1939) for an exhaustive historical review from the viewpoint of stellar formation. Interest in the Debye-Hückel system was recently revived, when the “Nernst-Planck” (NP) equations in the same form as (1) re-emerged in the biophysics community, for example, in the study of ion transport in biological channels. In this case, the flux $`𝐉`$ is called the Nernst-Planck (NP) particle flux for the ion current density $`𝐉`$ depending linearly on the gradients $`(\rho ,\mathrm{\Phi })`$ and $`L`$ is the Laplace operator. See Barcilon et al. (1992) and Syganow and von Kitzing (1999) for recent reviews of ion transport in biological media. The same elliptic-parabolic system had also surfaced earlier as the “drift-diffusion” equations in the semiconductor device design literature, Selberherr (1984) . Similar elliptic-parabolic equations (with hyper-diffusion, instead of diffusion) also arose recently as a leading order description of molecular beam epitaxy . A variant of the system (1) re-appeared even more recently as a model of self-assembly of particles at nano-scales . ## 2 Derivation of aggregation equation for <br>finite-size particles and variable mobility In this paper, we modify the class of Debye-Hückel equations (1) with $`ϵ=0`$, as a model of the aggregation of interacting particles of finite size. This problem is motivated by recent experiments using self-assembly of nano-particles in the construction of nano-scale devices -. A colloidal solution of 50nm-size particles is deposited on a grooved substrate; evaporation and receding contact line drags the particles into the channels. It has been shown that the most important phenomenon effecting the distribution of particles in the channels is the capillary interaction between the particles at last stage of water evaporation, when all water and particles are confined to the channels. Examples of particle self-assembly in nano-channels is shown inFig. 1. Fundamental principles underlying the mathematics of self-assembly at the nano-scales are non-local particle interaction and nonlinear motion due to variations of mobility at these scales. Let us start by deriving the effective interaction potential between nano-particles. This will reveal the length scales involved in the problem and characterize the type of interaction between the particles. Due to the effects of surface tension, the air/water interface is deformed near a particle. On the other hand, water is attracted to the substrate surface by van der Waals forces. Denote the thickness of the undisturbed water layer by $`H_0`$ and let $`h`$ be the elevation from this equilibrium level. If we call the potential of van der Waals interaction $`U_{vdW}(H)`$, then for small deviations from the equilibrium surface the balance of surface tension and van der Waals force gives $$\gamma \mathrm{\Delta }h=F_0h,F_0=\frac{U_{vdW}}{H}(H=H_0)$$ (2) A typical expression for van der Waals potential includes three terms $$U_{vdW}(H)=\frac{A}{H^3}\frac{B}{H^\kappa }Ce^{H/l_0}$$ (3) with constants $`A`$, $`B`$, $`C`$, with $`\kappa >3`$ and $`l_0`$ being the interaction length. A numerical estimate for these values from shows that for $`H50`$ nm, the first term on the right-hand side of (3) is larger than the second and third terms by factors $`10^9`$ and $`10^{18}`$, respectively. Here, $`A`$ is called *Hamaker constant*, which assumes a certain value for a given pair of materials.<sup>2</sup><sup>2</sup>2Usually, the numerical vallue of $`A`$ is given in *zJoules*, with *z* being used for *zepto*$`=10^{21}`$. Then, (2) can be re-written in terms of *capillary length* $`l_c`$ as $$l_c^2\mathrm{\Delta }hh=0l_c=H_0^2\sqrt{\frac{\sigma }{3A}}$$ (4) where we have applied the term *capillary length* even though the stabilizing potential is due to van der Waals forces and not gravity. Numerical values for Hamaker constant for water on silicon is about $`30zJ=3010^{21}J`$, surface tension of water/air interface is $`0.07N/m`$. Consequently, for $`H_0=50`$nm, equation (4) yields capillary length $`l_c300`$nm, or several times the particle diameter. It is now clear that in spite of very different physics and length scales, the phenomenon of self-attraction of particles at nano-scales bears not just qualitantive, but also quantitative resemblance to the well-known *Cheerios effect*: self-attraction of floating bodies due to surface tension. For floating cheerios, the typical length of interaction is the usual capillary length (several mm), which is also of the order of particle size. It has been shown that two particles separated by a distance $`l`$ have interaction potential proportional to $`\mathrm{exp}(l/l_c)`$ in one dimension, and $`K_0(l/l_c)`$ in two dimensions ($`K_0(x)`$ being the modified Bessel function of the first kind). The force between the particles is then proportional to $`K_1(l/l_c)`$, and this result holds for both floating and partially submerged particles. Mathematically, these interaction potentials arise from inversion of Helmholtz operator in (4). These physical considerations show that interaction potentials proportional to the Green’s function for the Helmholtz operator plays a fundamental role in particle self-assembly across surprisingly many orders of magnitude. Thus, we shall consistently use this interaction potential in all our numerical simulations. Nevertheless, we shall keep our framework general enough that our method of modeling nonlocal interactions among different particles may be extended to the other physical problems of interest described in the previous section. Our method takes into account the change of mobility due to the finite size of particles and the nonlocal interaction among the particles. The local density (concentration) of particles is denoted by $`\rho `$. Suppose the particles interact pairwise via the potential $`G(|𝐫|)`$. The total potential at a point $`𝐫`$ is $$\mathrm{\Phi }(𝐫)=\rho (𝐫^{})G(|𝐫𝐫^{})\text{d}𝐫^{}=G\rho $$ (5) where $``$ denotes convolution. (The minus sign is chosen so $`G>0`$ for attracting particles.) The velocity of the particle is assumed to be proportional to the gradient of the potential times the mobility of a particle, $`\mu `$. The mobility can be computed explicitly for a single particle moving in an infinite fluid. However, when several particles are present, especially in highly dense states, the mobility may be hampered by the particle interactions. These considerations are confirmed, for example, by the observation that the viscosity of a dense suspension of hard spheres in water diverges, when the density of spheres tends to its maximum value. Many authors have tried to incorporate the dependence of mobility on local density by putting $`\mu =\mu (\rho )`$ . It is common to assume that the mobility is a function of density, which tends to zero when the density tends to some maximum value, assumed to be $`\rho _{\text{max}}=1`$, i.e., $`\mu (\rho _{\text{max}})=0`$. Vanishing mobility leads to the appearance of weak solutions in the equations, to singularities and, in general, to massive complications and difficulties in both theoretical analysis and computational simulations of the equations. In contrast, the Keller-Segel model specifies $`\mu (\mathrm{\Phi })`$, so the mobility is taken in that case as a function of the concentration of chemotactic agent (potential) $`\mathrm{\Phi }`$. The goal of this paper is to suggest and investigate the implications of the idea that the mobility $`\mu `$ should depend on the *averaged* density $`\overline{\rho }`$, rather than either the potential, or the exact value of the density at a point. This assumption makes sense from the viewpoints of both physics and mathematics. From the physical point of view, the mobility of a finite-size particle must depend on the configuration of particles in its vicinity. While attempts have been made to approximate this dependence by using derivatives of the local density, it is much more natural, in our opinion, to assume that the local mobility depends on an integral quantity $`\overline{\rho }`$ which is computed from the density as $`\overline{\rho }=H\rho `$. Here $`H(𝐫)`$ is some *filter* function having, in general, short range compared to the potential $`G`$. Several filter functions are possible, with examples being $`H(𝐫)=\delta (𝐫)`$ ($`\delta `$-function), $`H(𝐫)=\mathrm{exp}(|𝐫|/l)/(2l)`$ (exponential, or inverse-Helmholtz in 1D) or, in $`d`$ dimensions, $`H(𝐫)=\theta (|𝐫|l)/(2l)^d`$. The last is the top-hat function, whose value is unity when $`|𝐫|`$ is between $`l`$ and $`l`$, and vanishes, elsewhere. The normalizing factor of $`(2l)^d`$ ($`d`$ being the dimension of space) is introduced so that $`_{\mathrm{}}^+\mathrm{}H(𝐫)\text{d}𝐫=1`$. Alternatively, one may assume a filter function $`H(𝐫)`$ with zero average: $`_{\mathrm{}}^+\mathrm{}H(𝐫)\text{d}𝐫=0`$. The latter assumption may be useful in crystal growth models, where the mobility does not depend on the absolute level of the material. Rather, in such models the mobility depends on the relative positioning of particles. The mathematical analysis and the reduction property derived in this paper hold, regardless of the shape of the filter function $`H`$ and the potential $`G`$ is, so long as they are both *nice* (*e.g.*, piecewise smooth) functions. #### Aim of the paper The object of this paper is the analysis of the following continuity equation for density evolution, $$\underset{\text{Continuity equation}}{\underset{}{\frac{\rho }{t}=\mathrm{div}𝐉}}\text{with}\underset{\text{Particle flux}}{\underset{}{𝐉=D\overline{\rho }\mu (\overline{\rho })\rho \mathrm{\Phi }}}.$$ (6) Here $`D`$ is a constant diffusivity, while $`\overline{\rho }`$ and $`\mathrm{\Phi }`$ are defined by two convolutions involving, respectively, the filter function $`H`$ and the potential $`G`$, $$\overline{\rho }=H\rho \text{and}\mathrm{\Phi }=G\rho .$$ (7) While this equation preserves the sign of $`\rho `$, we shall see that it allows the formation of $`\delta `$-function singularities even when $`D>0`$ and in the case of one spatial dimension. Alternatively, system (6,7) with $`D=0`$ can be represented geometrically as $$\frac{d}{dt}(\rho \mathrm{dVol})=0\text{along}\frac{d𝐱}{dt}=𝐉(\rho ),$$ where the vector field corresponding to the current $`𝐉(\rho )`$ is defined in terms of the pointwise density $`\rho 0`$ and two prescribed functionals of $`\rho `$: an energy $`E(\rho )`$ and a mobility $`\mu (\overline{\rho })0`$, as $$𝐉(\rho )=\rho \mu (\overline{\rho })\frac{\delta E}{\delta \rho }.$$ Here $`\overline{\rho }=H\rho 0`$ is an average density. As a result, $$\frac{d}{dt}E(\rho )=\frac{1}{\rho \mu (\overline{\rho })}\left|𝐉(\rho )\right|^2\mathrm{dVol}0.$$ Thus, the flow $`d𝐱/dt=𝐉`$ causes the energy functional $`E(\rho )`$ to decrease toward its minimum value and the flow is Lyapunov stable, provided the energy functional $`E(\rho )`$ is sign definite and $`\mu (\overline{\rho })`$ is isolated from $`0`$. We may regard $`𝐉`$ as a vector field defining an infinitesimal action of the diffeomorphisms on the space of positive maps $`\rho `$ acting on a manifold $``$. Remarkably, this action produces singularities in which the pointwise density concentrates in finite time on subspaces embedded in the manifold $``$. #### Subcases The generality and predictive power of the model (6,7) can be demonstrated by enumerating a few of its subcases, as follows. When $`H(𝐫)=\delta (𝐫)`$ and $`G(x)=e^{|𝐫|/l}`$ the system reduces to the generalized chemotaxis equation . For the choice $`G=\delta ^{}(𝐫)`$, $`H=\delta ^{}(𝐫)`$ one obtains a modification of the inviscid Villain model for MBE evolution (with extra factor of $`\rho `$ in the flux). If the range of $`G`$ (denoted by $`l`$) is sufficiently small, we can approximate (at least formally) the integral operator in $`\mathrm{\Phi }=G\rho `$ as a differential operator acting on the density $`\rho `$, namely, $$\mathrm{\Phi }=\rho +l(l\rho ).$$ As was demonstrated in , equation (6) then becomes a generalization of the viscous Cahn-Hilliard equation, describing aggregation of domains of different alloys. In the case that $`\mu `$ is a constant, $`H(xy)=\delta (xy),`$ and $`G`$ is the Poisson kernel, then equation (6) becomes the drift limit of the Poisson-Smoluchowski equation for the interaction of gravitationally attracting particles under Brownian motion . All the particular cases described above require a singular choice of the functions $`G`$ and $`H`$. However, the generalized functions required in these cases may be approximated with arbitrary accuracy by using sequences of nice (for example, piecewise smooth) functions. We shall concentrate on cases where the functions $`H`$ and $`G`$ remain nice, and derive the results in this more general, more regular, setting. Thus, regularization in this endeavor introduces additional generality, which enables analytical progress and makes numerical solution of the equations easier. ## 3 Evolution of density as mass conserving gradient flow: the nonlocal Darcy’s law and energetic considerations Let us have another look at our motivation of equations (6,7), this time making a connection with gradient flows and thermodynamics. We will show that there is a naturally defined free energy which remains finite, even when the solutions cease to be smooth and are replaced by a set of delta-functions. We show that the energy remains finite for all choices of $`\mu (\overline{\rho })`$. In contrast, the traditional approach considering the dependence of mobility on the un-smoothed density $`\mu (\rho )`$ would fail by allowing the energy to become infinite. This additional energetic argument reinforces the choice of the regularized model in (6,7). The calculation will be performed in arbitrary spatial dimensions. For the energy to be well-defined, we require that the function $`G`$ describing interaction between two particles is everywhere positive, so the interaction is always attractive. In one dimension, we also require that the kernel function $`G`$ is symmetric and in $`n`$ dimensions that the interaction is central. This assumption is physically viable for all the historical cases described in the introduction. Our derivation for the free energy will remain valid for an *arbitrary* filter function $`H`$. We start with equation (6) for mass conservation, in the case when the mobility $`\mu (\overline{\rho })`$ in the particle flux $`𝐉`$ is not constant. We shall derive a variant of equation (6) as a gradient flow in the sense that, $`{\displaystyle \frac{\rho }{t}}=\mathrm{grad}E|_\rho ,\text{or}{\displaystyle \frac{\rho }{t}},\varphi ={\displaystyle \frac{\delta E}{\delta \rho }},\varphi ,`$ where $`f,\varphi =f\varphi d^nx`$ is the $`L^2`$ pairing, for a suitable test function $`\varphi `$ and where $`E`$ is the following energy, $$E[\rho ]=D\rho (\mathrm{log}\rho 1)\text{d}x+\frac{1}{2}\rho \mathrm{\Phi }\text{d}x\text{with}\mathrm{\Phi }=G\rho .$$ (8) The first term in this expression for the energy $`E[\rho ]`$ decreases monotonically in time for linear diffusion. The second term defines the $`H^1`$ norm for the choice $`G(x)=\mathrm{exp}(|x|/\alpha )`$. It may be regarded as a generalization of this norm for an arbitrary (but positive and symmetric) function $`G`$. Such negative Sobolev norms remain finite, even when the solution for the density $`\rho `$ concentrates into a set of delta functions. The variation $`\delta E/\delta \rho `$ of the free energy $`E[\rho ]`$ in equation (8) is $`\delta E[\rho ]={\displaystyle (D\mathrm{log}\rho +\mathrm{\Phi })\delta \rho d^nx},`$ where we understand $`\delta \rho `$ as arising from the flow of a diffeomorphism, as for example in . Namely, we specify $`\delta \rho =\mathrm{div}\left(\rho \mu (\overline{\rho })\mathrm{\Psi }\right),`$ for an arbitrary variation of $`\rho `$ determined by the function $`\mathrm{\Psi }`$, which is assumed to be smooth.<sup>3</sup><sup>3</sup>3This specification of the density variation formally requires the solution to remain differentiable and the product $`\rho \mu (\overline{\rho })`$ not to vanish. In principle, these conditions would be violated by the formation of weak solutions, in which the density concentrates into delta-functions. However, we will verify a posteriori that the gradient flow properties of the smooth solutions discussed in this section are also preserved for the weak solutions. The corresponding variational derivative is given by, cf. $`\delta E[\rho ]`$ $`=`$ $`{\displaystyle \frac{\delta E}{\delta \rho }},\mathrm{\Psi }={\displaystyle (D\mathrm{log}\rho +\mathrm{\Phi })\mathrm{div}\left(\rho \mu (\overline{\rho })\mathrm{\Psi }\right)d^nx}`$ (9) $`=`$ $`{\displaystyle \mathrm{\Psi }\mathrm{div}\left(\mu (\overline{\rho })\left(D\rho +\rho \mathrm{\Phi }\right)\right)d^nx}.`$ Consequently, the free energy $`E[\rho ]`$ in equation (8) produces the following gradient flow: $`{\displaystyle \frac{\rho }{t}},\mathrm{\Psi }={\displaystyle \frac{\delta E}{\delta \rho }},\mathrm{\Psi }=\mathrm{div}\left(\mu (\overline{\rho })\left(\mathrm{D}\rho +\rho \mathrm{\Phi }\right)\right),\mathrm{\Psi }.`$ The resulting variant of equation (6) is $$\underset{\text{Continuity eqn}}{\underset{}{\frac{\rho }{t}=\mathrm{div}𝐉}}\text{with}\underset{\text{Modified particle flux}}{\underset{}{𝐉=\mu (\overline{\rho })(D\rho +\rho \mathrm{\Phi }))}},$$ (10) in which the linear diffusivity of density $`\rho `$ is modified by the nonlocal mobility $`\mu (\overline{\rho })`$. #### Energetics The free energy $`E[\rho ]`$ in (8) decreases monotonically in time under the evolution equation (10). A direct calculation yields, $$\frac{dE[\rho ]}{dt}=(D\mathrm{log}\rho +\mathrm{\Phi })\mathrm{div}𝐉d^nx=\frac{1}{\rho \mu _(\overline{\rho })}|𝐉|^2d^nx.$$ (11) According to this equation, provided $`\rho \mu _(\overline{\rho })>0`$, the rate of decay of free energy $`E[\rho ]`$ given in (8) defines a Riemannian metric in the particle flux, cf. . We shall see that when $`D=0`$, the resulting conservative motion of finite-size particles drifting along the gradient of $`\mathrm{\Phi }`$ leads to a type of ‘clumping’ of the density $`\rho `$ into a set of delta functions. One may check that this monotonic decrease of energy persists for more general functions, including weak solutions supported on $`\delta `$-functions, as discussed below. ## 4 Non-vanishing mobility: weak solutions <br>(*clumpons*) and their role in long-term dynamics in 1D Everywhere in this paper, we shall assume that the potential is purely attractive, so $`G(x)>0`$ for all $`x`$. This assumption is used for two main reasons. First, it suits the physics of the motivating problem, namely, mutual attraction of nano-particles . Second, restricting to purely attractive interactions allows one to separate all equations of the type (6,7) into two different classes. The physical property separating the two classes is whether the mobility $`\mu (\overline{\rho })`$ is strictly isolated from zero, i.e., $`\mu (\overline{\rho })\mu _0>0`$ or one may allow $`\mu (\overline{\rho }_0)=0`$ for some $`\overline{\rho }=\overline{\rho }_0`$. In the first case, there is nothing to resist the mutual attraction of the particles and the final state of the system is a single $`\delta `$-function no matter what the form of the mobility dependence, as long as mobility is strictly isolated from zero. In the second case, vanishing of the mobility at some maximum density blocks the motion when the densities become too large and prevents total collapse. Instead of collapse, this produces isolated *patches* of solutions of constant density. It is interesting that also in this second case, the stationary solutions remain *exactly the same* regardless of the dependence of $`\mu `$ on $`\overline{\rho }`$, and we can describe these stationary solutions analytically. Moreover, we shall demonstrate that these solutions are stable and any initial condition separates into a set of isolated stationary patches of this type. This section is devoted to analytical description of dynamics in the case when $`\mu (\overline{\rho })`$ being strictly isolated from zero. In simulations, we have assumed $`\mu (\overline{\rho })=1`$ for simplicity, but all our results will remain valid for arbitrary dependence of $`\mu `$ on $`\overline{\rho }`$, as long as $`\mu `$ always remains positive. ### 4.1 Formal weak solution anzatz This section considers motion under equations (6,7) in one spatial dimension for the case of vanishing linear diffusivity, $`D=0`$. Substituting the following singular solution ansatz $`\rho (x,t)`$ $`=`$ $`{\displaystyle \underset{i=1}{\overset{N}{}}}w_i(t)\delta (xq_i(t)),\overline{\rho }(x,t)={\displaystyle \underset{j=1}{\overset{N}{}}}w_j(t)H(xq_j(t)),`$ (12) into the one-dimensional version of equation (6) with $`D=0`$ and integrating the result against a smooth test function $`\varphi `$ yields $`{\displaystyle \varphi \left[\rho _t\left(\rho \mu (\overline{\rho })(G\rho )_x\right)_x\right]𝑑x}={\displaystyle \varphi (x)\underset{i=1}{\overset{N}{}}\dot{w}_i\delta (xq_i)dx}`$ $`+{\displaystyle \varphi ^{}(x)\underset{i=1}{\overset{N}{}}w_i\left(\dot{q}_i+\underset{j=1}{\overset{N}{}}w_j(t)\mu \left(\overline{\rho }\right)G^{}(xq_j)\right)\delta (xq_i)dx}`$ Hence, one obtains a closed set of equations for the parameters $`w_i(t)`$ and $`q_i(t)`$, $`i=1,2,\mathrm{},N,`$ of the solution ansatz (12), in the form $`\dot{w}_i(t)=0,\dot{q}_i(t)={\displaystyle \underset{j=1}{\overset{N}{}}}w_j\mu \left(\overline{\rho }\right)G^{}(q_iq_j)`$ (13) where $$\overline{\rho }=\underset{m=1}{\overset{N}{}}w_mH(q_m(t))$$ (14) Thus, the density weights $`w_i(t)=w_i(0)=w_i`$ are preserved, and the positions $`q_i(t)`$ follow the characteristics of the velocity $`𝐮=\mu (\overline{\rho })G\rho `$ along the Lagrangian trajectories at $`x=q_i(t)`$. This result holds in any number of dimensions, modulo changes to allow singular solutions supported along moving curves in 2D and moving surfaces in 3D. Fig. 2, demonstrates that the solutions (13) do indeed appear spontaneously in a numerical simulation of equation (6) with $`D=0.02`$. Hence, they are ubiquitous and dominate its dynamics. The simulation started with a smooth (Gaussian) initial condition for density $`\rho `$. Almost immediately, one observes the formation of several singular *clumpons*, which evolve to collapse eventually into a single clumpon. Observe that the mass of each individual clumpon remains almost exactly constant in the simulations, as required by equation (13). Also note that the masses of two individual clumpons add when they collide and “clump” together. Eventually, all the mass becomes concentrated into a single clumpon, whose mass (amplitude) is *exactly* the total mass of the initial condition. ### 4.2 Energy decay close to the final state and estimates for collapse time Normally, systems approaching an equilibrium state tend to evolve more slowly as they approach the equilibrium. If the rate of approach diminishes linearly with the distance from equilibrium (in some sense), for example, one obtains an exponential decay towards the equilibrium. Alternatively, for finite-time singularities, the rate tends to diverge to infinity in a power-law fashion. In contrast to these familiar examples, the system (6,7) approaches the singularity at a *constant* rate. That is, the rate of approach to singularity never diverges. Consequently, one may predict the formation of singularities and even predict their evolution after they have formed. This surprising result may be demonstrated by deriving an alternative form of energy dissipation. Direct substitution of a single $`\delta `$-function for density into (11) is mathematically ambiguous, as it would lead to improper operations with $`\delta `$-functions. Instead, let us notice that the evolution of the energy $`E`$ describing the gradient flow of (6,7) can be expressed as follows $$\frac{dE}{dt}=G(xx^{})\rho (x^{})\frac{\rho (x)}{t}\text{d}x\text{d}x^{}=$$ $$\begin{array}{cc}& +\mathrm{\Phi }\text{div}\left(\rho \mu (\overline{\rho })\text{grad}\mathrm{\Phi }\right)=\rho \mu (\overline{\rho })(\mathrm{\Phi })^2\hfill \end{array}$$ (15) Note the difference between this formula and (11). Formally, (15) is the same as (11), but we can substitute the delta-function ansatz for weak solutions directly into (15). Numerical simulations, energetic considerations and physical intuition suggest that the final state of the system with strictly positive mobility $`\mu (\rho )\mu _0>0`$ should be a single clumpon. One may ask how the time necessary to collapse to this final state varies with the initial conditions. An exact analytical answer to this question seems out of reach and we shall provide a numerical simulation for various initial conditions. However, a surprisingly simple and accurate analytical estimate can be made here. To start, let us compute the rate of energy dissipated by a single clumpon, which is the final state of our system. We assume $`\rho =M\delta (x)`$, where $`M`$ is the total mass. Assuming that $`H(x)`$ is regular, $`H\rho (x)=MH(x)`$. Let us also assume for simplicity that $`G(x)=G_0\mathrm{exp}(|x|/\alpha )`$. Direct computation gives $$\frac{dE}{dt}[\rho =\text{clumpon}]=M\mu \left(MH(0)\right)\left(\mathrm{\Phi }(0)\right)^2=\frac{M^3}{\alpha ^2}\mu \left(MH(0)\right)G_0^2$$ (16) We may now estimate how long it will take for an arbitrary system to collapse into a single clumpon. If the initial energy of the system is $`E_0`$, and initial mass is $`M`$, the energy of a single clumpon will be $$E_f=M^2G_0.$$ (17) Then, the approximate time to collapse to a single clumpon from any initial conditions is given by combining (16) and (17): $$t_{}\frac{E_0E_f}{dE/dt[\rho =\text{clumpon}]}=\frac{E_0M^2G_0}{M^3\mu \left(MH(0)\right)G_0^2}$$ (18) In derivation of (18) we have assumed that $`dE/dt`$ does not change much along the trajectories, so the initial conditions are close enough to a clumpon. Thus, this estimate for the time of collapse depends on only two integral quantities: initial energy and mass for initial conditions close to the final state. ### 4.3 Rate of blow up The following analysis of the evolution of a density maximum reveals that the clumping process results from the nonlinear instability of the gradient flow in equation (6) when $`D=0`$. For a particular case $`G=G_0e^{|x|/\alpha }`$ and $`\mu (\overline{\rho })=1`$, one may show that for a high enough peak, a density maximum $`\rho _m(t)=\rho (x_m(t),t)`$ becomes infinite in finite time. The motion of the maximum is governed by $$\frac{d}{dt}\rho _m=\frac{1}{\alpha ^2}\left(\rho _m^2\rho _m\mathrm{\Phi }(x_m)\right)\frac{1}{\alpha ^2}\left(\rho _m^2\rho _mM\right),$$ (19) where $`M=\rho \text{d}x`$ is total mass and we have used the fact that $`\mathrm{\Phi }`$ satisfies $`\mathrm{\Phi }\alpha ^2\mathrm{\Phi }_{xx}=\rho `$. The last inequality holds, because $`G1`$ is bounded and $`\rho `$ is everywhere positive. Thus, if at any point the maximum of $`\rho `$ exceeds the (scaled) value of the total mass, then the value of the density maximum $`\rho _m(t)`$ must diverge in finite time. This divergence produces $`\delta `$-functions in finite time. From (19), the density amplitude must diverge as $`\rho _m\alpha ^2/(t_0t)`$. The formation of singularities in Fig. 2 occurs both at the maximum, and elsewhere. The subsidiary peaks eventually collapse with the main peak. ### 4.4 Dynamics of inflection points and collapse Let us discuss in more detail the process by which density singularities are formed. Suppose the the initial condition contains only one density maximum. Then, one would expect the first singularity to form at the position of this maximum. As this density singularity forms, mass flows toward it. This causes a local reduction of density in the neighboring region from which the mass is flowing. Whenever this local reduction of density due to the formation of the singularity is sufficient to form a new maximum in density away from it, then the process of singularity formation starts again there, and so forth. In principle, this process could form an infinite number of density singularities. However, we propose two reasons why the expected number should be finite. First, the range of interaction is set by the length scale in $`H`$. Thus, one might expect no more clumpons to form than the ratio of the interaction range to the domain size. Second, in the formation of singularities from two nearby maxima, one of them may become stronger than the other and entrain it, thereby suppressing the formation of the second singularity. In practice, one sees the formation of considerably fewer clumpons than the number estimated by the ratio of the interaction range to the domain size. Hence, one may conjecture that the formation of singularities does involve some competition between neighboring density maxima. This conjecture could be tested by studying evolution from initial conditions containing several density maxima separated by varying distances which are comparable to, or smaller, than the average length scale in $`H`$. A quantitative evaluation of this heuristic argument may be obtained by computing the position of the inflection point for the averaged density $`\overline{\rho }`$. Although no analytical formula for the motion of inflection point is available, one sees numerically that the inflection point quickly converges to the position of the singularity. The evolution of the inflection point corresponding to the solution from Fig. 2 is shown in Fig. 3. ## 5 Formation of jammed states when mobility $`\mu `$ approaches zero ### 5.1 Competing length scales of $`H`$ & $`G`$ An interesting limiting case arises when the scale of non-locality of $`H`$ is much shorter than the range of the potential $`G`$. Formally, this limit corresponds to $`H(x)\delta (x)`$. In practice, we may select a sequence of piecewise smooth functions $`H_ϵ(x)=\mathrm{exp}(|x|/ϵ)/(2ϵ)`$ which converge weakly to a $`\delta `$-function. For each function $`H_ϵ(x)`$ in this sequence, no matter how small (but positive) the value of $`ϵ`$, the exact ODE reduction (13) still holds. So, what happens in the limit of very small epsilon, as $`ϵ0`$? In investigating this limit, we performed a sequence of numerical simulations for a fixed choice of the function $`G(x)=\mathrm{exp}(|x|/\alpha )`$ while varying $`H_ϵ(x)=\mathrm{exp}(x/ϵ)/(2ϵ)`$ over a sequence of decreasing values of the ratio $`ϵ/\alpha `$. This simulation is shown in Fig. 2 for $`ϵ/\alpha =1/10`$ and it demonstrates the formation of flat clumps of solutions. The mechanism for this phenomenon is the following. The vanishing mobility at $`\overline{\rho }=1`$ caps the maximal density at $`\overline{\rho }=1`$ in the long-term. This leads to the appearance of flat mesas in $`\overline{\rho }(x)`$ for large $`t`$. On the other hand, when $`H(x)\delta (x)`$, one finds $`\overline{\rho }(x,t)\rho (x,t)`$ pointwise in $`x`$, which forces $`\rho (x,t)`$ to develop a flat mesa, or plateau, structure in which the maximum is very close to unity, as well. This is precisely what is predicted for the chemotaxis equation with $`H(x)=\delta (x)`$ and $`\mu (\rho )=1\rho `$ . In the limit $`H\delta `$, model (6,7) recovers ordinary diffusion of local density. This is a singular limit, because it increases the order of the differentiation in the equation. Since ordinary diffusion is known to prevent collapse in one dimension , this singular limit should be of considerable interest for further analysis. ### 5.2 Stationary states Numerical simulations of time-dependent problem (6,7) show that the solution converges to well-defined states with flat *mesa* peaks when $`\mu (\overline{\rho })`$ reaches $`0`$ for some value of $`\overline{\rho }`$. In the remainder of this section, we shall concentrate on the analytical description of the evolution in this case. For simplicity of formulas, we shall assume that the critical value of $`\overline{\rho }`$ is normalized to be $`1`$, *i.e*., $`\mu (1)=0`$. In simulations, we shall take $`\mu =1\overline{\rho }`$. However, all theoretical results, in particular, exact expressions for stationary sates, remain true for *arbitrary* dependence of $`\mu (\overline{\rho })`$, as long as $`\mu (\overline{\rho })`$ reaches $`0`$ at some value of $`\overline{\rho }`$. We distinguish two classes of stationary states, which differ mathematically and physically. Both classes of stationary states will describe a clump of particles whose density $`\rho (x)`$ is a generalized function with compact support, which we will call a *patch*. The difference between these classes lies in the physical reason for the vanishing of local velocity. Remember that local velocity is written as $`𝐮=\mu (\overline{\rho })\mathrm{\Phi }`$. The solution is stationary if $`𝐮`$ vanishes at every point inside the patch. This can be achieved in two ways. First, $`\mathrm{\Phi }`$ may be exactly zero everywhere inside the patch. In this case, the net force acting on every particle vanishes. These solutions are called *equilibrium* states. Second, it may also happen that the mobility $`\mu `$ vanishes at every point inside the patch. Physically, this corresponds to a clump at maximum density whose motion is prohibited, although the net force on each particle may not be zero. These solutions are called *jammed* states. Let us now study both the equilibrium and jammed states in more details. As we shall see, an analytical solution describing each of these states can be found. These solutions will also elucidate the nature of competing length scales in $`G`$ and $`H`$. The stability of these solutions and their selection mechanisms will also also be demonstrated. ### 5.3 Equilibrium states Let us assume for now that both the potential attraction $`G(x)`$ and the filter function $`H(x)`$ are proportional to the Green’s function for the Helmholtz operator. This is both the case we use in numerical simulations, and also, incidentally, the only case we have found that admits analytical solution for stationary states. Thus, we posit $$G(x)=\mathrm{exp}\left(\frac{|x|}{\alpha }\right)H(x)=\frac{1}{2\beta }\mathrm{exp}\left(\frac{|x|}{\beta }\right)$$ (20) We are looking for stationary states $`\rho (x)`$ that are weak solutions with compact support, for which $`\rho (x)=0`$ if $`|x|>L`$. While it is impossible to find a smooth (or even piecewise smooth) function $`\rho (x)`$ which is a stationary solution of (6), one can find a stationary solution of the following form $$\rho (x)=w\delta (x+L)+w\delta (xL)+\{\begin{array}{cc}0,\hfill & |x|>L\\ \rho _0(x),\hfill & |x|L,\end{array}$$ (21) where the value of the constants $`w`$ and $`L`$ are to be determined. For (21) to be a stationary (weak) solution, the following conditions must be satisfied: $$\begin{array}{cc}\mathrm{\Phi }(x)=(G\rho )(x)=\text{const},& |x|<L\\ \overline{\rho }(x)=(H\rho )(x)=1,& x=\pm L.\end{array}$$ (22) Substitution of (21) into the first condition of (22) gives $$\mathrm{\Phi }(x)=wG(xL)+wG(x+L)+_L^LG(xy)\rho _0(y)\text{d}y=\text{const}$$ (23) It is crucial to notice that if $`G(x)`$ is given by (20), then the integral in (23) must yield a function proportional to a sum of exponentials, which is only possible if $`\rho _0(x)`$ is a constant: $`\rho _0(x)=\rho _0`$. Performing the integral gives $$\mathrm{\Phi }(x)=we^{|xL|/\alpha }+we^{|x+L|/\alpha }+\rho _0\alpha \left(2e^{|xL|/\alpha }e^{|x+L|/\alpha }\right)=\text{const}$$ (24) which requires $$w=\rho _0\alpha .$$ (25) Enforcing the second condition in (22) will determine $`\rho _0`$ as a function of $`L`$. As in (23), we find an analytical expression for $`\overline{\rho }`$ $$\overline{\rho }(x)=\frac{w}{2\beta }\left(e^{|xL|/\beta }+e^{|x+L|/\beta }\right)+\frac{\rho _0}{2}\left(2e^{|xL|/\beta }e^{|x+L|/\beta }\right)$$ (26) Since (26) assumes identical values at $`x=\pm L`$, we need only check that $`\overline{\rho }(x=L)=1`$. Substitution of $`x=L`$ into (26) gives $$\rho _0\left[\alpha \left(1+\mathrm{exp}\left(2L/\beta \right)\right)+\beta \left(1\mathrm{exp}\left(2L/\beta \right)\right)\right]=2\beta $$ (27) In principle, equation (27) already determines the density inside the patch as a function of length. It is, however, more practical to determine the length as a function of total mass of the clumpon $`M`$. The mass contained in the solution (21) is $$M=\rho _0L+2w=\rho _0(L+2\alpha ).$$ Thus, we find an implicit condition for length $`L`$ as a function of patch mass $`M`$: $$M\left[\alpha \left(1+\mathrm{exp}\left(2L/\beta \right)\right)+\beta \left(1\mathrm{exp}\left(2L/\beta \right)\right)\right]=2\beta (L+2\alpha )$$ (28) These solutions are shown on Fig. 5. It is interesting to note that (28) defines a non-negative length $`L`$ only if the mass $`M`$ is sufficiently large. This is perfectly physical: if the mass is small, there is no reason for the solution to “spread out”, so a single $`\delta `$-function is produced. Incidentally, this will happen if the mass of the clumpon $`M=\rho \text{d}x`$ is such that $`max_x\overline{\rho }<1`$, i.e., $`M<2\beta `$. Physically, we expect that equilibrium solutions are unstable. Indeed, imagine a set of particles on a line positioned at an equal distance from each other. If the force between each pair of particles is repulsive, this configuration is stable, and the real part of all eigenvalues of the linearized problem is negative. However, if the force between the particles changes to purely attractive without changing the particle positions, this leads to the sign change of the eigenvalues which correspond to a stable situation. Thus, these equilibrium states with fixed gaps between the particles “held together” by purely attractive forces must be unstable, and must exhibit large values for the real parts of the unstable eigenvalues. Numerical analysis of linear stability of equilibrium states in the continuum description confirms this intuitive physical picture. Namely, there exists a set of eigenvalues with large positive real parts. In addition, a fully nonlinear time-dependent simulation of (6,7) starting with initial conditions corresponding to even slightly perturbed equilibrium states shows very rapid deviation away from equilibrium. Therefore, equilibrium states are *unstable* and will never be realized in nature. ### 5.4 Jammed states The derivation of the jammed states is rather similar to that for the equilibrium states, so only a brief discussion of it will be provided. Let us again assume that stationary states $`\rho (x)`$ are weak solutions with compact support, and $`\rho (x)=0`$ if $`|x|>L`$: $$\rho (x)=w\delta (x+L)+w\delta (xL)+\{\begin{array}{cc}0,\hfill & |x|>L\\ \rho _0(x),\hfill & |x|L\end{array}$$ (29) Now, for solution (29) to be a jammed solution, we need just one condition: $$\overline{\rho }(x)=(H\rho )(x)=1,|x|\pm L.$$ (30) Substituting (29) into condition (30) gives $$\overline{\rho }(x)=wH(xL)+wH(x+L)+_L^LH(xy)\rho _0(y)\text{d}y=1$$ (31) Again, it is essential that $`H(x)`$ is given by (20), so the integral in (23) must be proportional to a sum of exponentials, which is only possible if $`\rho _0(x)`$ is a constant: $`\rho _0(x)=\rho _0`$. Equation (31) transforms to $$\overline{\rho }(x)=\frac{w}{2\beta }\left(e^{|xL|/\beta }+e^{|x+L|/\beta }\right)+\frac{\rho _0}{2}\left(2e^{|xL|/\beta }e^{|x+L|/\beta }\right)=1$$ (32) which requires $$w=\beta \rho _0=1$$ (33) The mass contained in the solution (21) is $$M=\rho _0L+2w=L+2\beta .$$ The jammed states are illustrated on Fig. 6. To understand the nonlinear stability of jammed states, let us again appeal to the physical picture of particles on a line. *Jammed* states correspond to a set of finite-size particles pressed tightly together. Physical intuition tells us that such a state should be favorable for purely attractive forces between pairs of particles. Numerical analysis of linear stability of *jammed* states in the continuum approximation confirms this: the real part of the spectrum is isolated from $`0`$ by $`D`$, where $`\lambda =D`$ is the limiting point of spectral sequence. Thus, $`D`$ should determine the time scale for convergence to stationary solutions. That is, the rate of convergence to stationary solution is given by the time scale $`\tau 1/D`$. To verify these predictions, a fully nonlinear simulation starting with a smooth initial conditions in the form of Gaussian peak has been performed. The results of this simulation are given in Fig. 7. For $`D=0.01`$ used in the simulation, the solution becomes practically stationary after $`t100`$. The small “bumps” at the position of the $`\delta `$-functions in density correspond to a mismatch in the stationary solutions, which were derived for $`D=0`$. The amplitudes of the “bumps” tend to zero when $`D`$ becomes very small. ## 6 Jammed stationary states in two dimensions ### 6.1 Exact jammed states in two dimensions While this paper focuses primarily on the evolution of particles in one dimension, the real-world technological importance of self-assembly processes requires us to derive and analyze stationary solutions of our model in two dimensions. Numerical simulations of particle dynamics show that there is a tendency for the formation of isolated clumps of either roughly circular shape, or, for high particle densities, roughly circular or elliptical voids in fully dense areas. To model these results, let us assume that the mobility $`\mu (\overline{\rho })`$ vanishes at critical density $`\overline{\rho }=1`$, and $`\rho `$ and $`\overline{\rho }`$ are connected through the two-dimensional Helmholtz operator $$\overline{\rho }\beta ^2\mathrm{\Delta }\overline{\rho }=\rho ,\text{so that}\overline{\rho }=H\rho .$$ (34) Inspired by these results we use the intuition developed in one dimension to seek jammed solutions in two dimensions which are fully dense $`\overline{\rho }=1`$ inside an area $`D`$ with additional distributed $`\delta `$-function for density on the boundary $`D`$, and $`\rho =0`$ in the exterior of $`D`$. Remarkably, when $`H`$ is the Green’s function for the Helmholtz operator, an exact analytical expression for several possible shapes of $`D`$ can be found. Again, we shall emphasize that our results will hold for an *arbitrary* functional dependence $`\mu (\overline{\rho })`$ as long as the mobility vanishes for some value of $`\overline{\rho }=\overline{\rho }_{}`$. For convenience, we have again rescaled the critical value to be $`\overline{\rho }_{}=1`$. Consider an orthogonal coordinate system $`(\xi ,\eta )`$ for which the solution of Helmholtz equation separates variables. Suppose, in addition, that the boundary of $`D`$ corresponds to one of the coordinate lines $`\xi =\xi _0`$, and interior of $`D`$ is given by $`\xi <\xi _0`$ Then, we may seek the solution in the exterior of $`D`$ in the form $`\overline{\rho }(\xi ,\eta )=F(\xi /\beta )G(\eta /\beta )`$. Boundary conditions for the exterior of $`D`$ are $`\overline{\rho }0`$ as $`\xi +\mathrm{}`$, and boundary conditions at $`\xi =\xi _0`$ are obtained by continuity of $`\overline{\rho }`$ at the boundary. Since for jammed states $`\overline{\rho }=1`$ in the interior, the continuity condition gives $`\overline{\rho }=1`$ on the boundary. Therefore, separable solutions must take the form $`\overline{\rho }(\xi ,\eta )=F(\xi /\beta )/F(\xi _0/\beta )`$. Since the system of coordinates is orthogonal and $`\rho =1`$ in the interior of $`D`$ from (34), we can compute the amplitude of $`\delta `$-function for density $`\rho `$ at the boundary $`\xi =\xi _0`$, which comes out to be $`\beta F^{}(\xi _0/\beta )/F(\xi _0/\beta )`$. Thus, the exact solution for the case when separation of variables of Helmholtz equation is possible and the boundary is given by $`D=\left\{\xi =\xi _0\right\}`$ is $$\rho (\xi ,\eta )=\{\begin{array}{cc}1,\hfill & (\xi ,\eta )\text{int}D\\ \delta (\xi \xi _0)\beta F^{}(\xi _0/\beta )/F(\xi _0/\beta )\hfill & (\xi ,\eta )D\\ 0,\hfill & (\xi ,\eta )\text{ext}D\end{array}$$ (35) $$\overline{\rho }(\xi ,\eta )=\{\begin{array}{cc}1,\hfill & (\xi ,\eta )\text{int}D\\ F(\xi /\beta )/F(\xi _0/\beta ),\hfill & (\xi ,\eta )\text{ext}D\end{array}$$ (36) To be specific, in Table 1 we enumerate all cases for which separation of variables in Laplace/Helmholtz equation is possible. As it is known, Helmholtz equation in two dimensions is separable in four coordinates only: Cartesian (which we shall not consider here), cylindrical, elliptic cylindrical and parabolic cylindrical . Each of these cases selects a particular shape of the patch, as well as a particular form of the function $`F(\xi )`$. Of course, similar results may be obtained in three dimensions, where eleven coordinate systems exist for which the Helmholtz differential equation is separable. We shall not go into details here as, first of all, the generalization is straightforward, and, second, we are only interested here in one- and two-dimensional self-assembly. Three-dimensional analogues of self-assembly may be important for some models arising in mathematical biology which are discussed at the end of Sec.7. ### 6.2 Irregular shapes: asymptotic result For arbitrary shape of domain $`D`$, it is no longer possible to find an exact expression for densities $`\rho `$ and $`\overline{\rho }`$. However, we may find an *asymptotic* result which shows that $`\delta `$-function density on the boundary must be a slowly-varying function of the boundary coordinates in the case when $`\beta `$ – the range of $`H`$ – is small compared to the size of the patch. We consider a patch of an arbitrary simple-connected shape $`\mathrm{\Omega }`$ with a smooth boundary and introduce the coordinate $`\eta `$ along the boundary of the patch $`\mathrm{\Omega }`$. Considerations of one dimensional case indicate that the patch is stable if it is jammed, i.e. $`\mu =0`$ everywhere inside the patch, which leads to equation $`\overline{\rho }=1`$ inside the patch. For now, we consider $`H(x)`$ to be inverse Hemholtz operator, thus $`\overline{\rho }`$ satisfies (34). Thus, inside the patch, we necessarily have $`\rho =1`$. The decay of $`\overline{\rho }`$ close to the boundary should be compensated by the presence of delta-function concentration at the boundary. Let us assume that the boundary is smooth. If $`\xi `$ is the coordinate locally perpendicular to the boundary and the boundary is at $`\xi =\xi _0`$, then we assume that the density near the boundary has the form: $$\rho (\xi \xi _0)=f(\eta )\delta (\xi \xi _0)+\text{ind}_\mathrm{\Omega }(𝐫),$$ (37) where $`\text{ind}_\mathrm{\Omega }(𝐫)`$ is the indicator function which equals unity. if $`𝐫\mathrm{\Omega }`$ and vanishes otherwise. From the condition $`\overline{\rho }=1`$ and $`\rho =1`$ in the interior of $`\mathrm{\Omega }`$ we obtain an integral equation for the strength $`f(s)`$: $$_\mathrm{\Omega }f(s)H(𝐫𝐫(s^{}))\text{d}s^{}+_\mathrm{\Omega }H(𝐫𝐫^{})\text{d}𝐫^{}=1$$ (38) If $`𝐫`$ is farther than $`\beta `$ away from the boundary, then only the contribution from the second integral is relevant. However, due to the extremely short range of $`H`$, this integral is equal to unity and equation (38) gives an identity. Thus, we only need to investigate equation (38) in the case when the distance between $`𝐫`$ and the boundary is of order $`\beta `$ or less. Suppose this distance is $`d\beta `$ with $`d>0`$ being a constant of order unity or less. Since $`H(𝐫^{}𝐫)`$ decays rapidly away from $`𝐫`$ for distances larger than $`\beta `$, we can approximate the slowly varying function $`f(s^{})`$ by its value at $`s`$, and the integral (38) as $$f(s)_\mathrm{\Omega }H(𝐫𝐫(s^{}))\text{d}s^{}+_\mathrm{\Omega }H(𝐫𝐫^{})\text{d}𝐫^{}=1$$ (39) Since we are only interested in the immediate neighborhood of the point $`𝐫(\eta )`$, we will now assume that the boundary is locally straight and vertical, and the interior of the patch is to the right of the boundary. We introduce the local coordinates $`x=\beta \xi `$ and $`y=\beta \eta `$ centered at the point $`𝐫`$, so the boundary is at $`x=d`$ and the $`x`$axis is pointing towards the boundary. Note the exact form of the kernel $`H`$ $$H(𝐫)=\frac{1}{2\pi \beta ^2}K_0\left(\frac{|𝐫|}{\beta }\right).$$ (40) We split up the second integral in (39) and perform the integration exactly by going to polar coordinates (and skipping algebraic details): $$_\mathrm{\Omega }H(𝐫𝐫^{})\text{d}𝐫^{}\frac{1}{2\pi }_{x<d}K_0\left(\sqrt{x^2+y^2}\right)\text{d}x\text{d}y=$$ (41) $$=1\frac{1}{2\pi }_{x>d}K_0\left(\sqrt{x^2+y^2}\right)\text{d}x\text{d}y=1\frac{1}{2}e^d$$ Let us now compute the first integral in (39). We have: $$H(𝐫(s)𝐫(s^{}))\text{d}s^{}=\frac{1}{2\pi \beta }_{y=\mathrm{}}^+\mathrm{}K_0\left(\sqrt{d^2+y^2}\right)\text{d}y=\frac{1}{2\beta }e^d$$ Thus, equation (39) becomes $$f(\eta )\frac{1}{2\beta }e^{|𝐫𝐫^{}|/\beta }+1\frac{1}{2}e^{|𝐫𝐫^{}|/\beta }=1$$ (42) so the answer is simply $$f(\eta )=\beta .$$ (43) Thus, for a smooth boundary, jammed states are obtained by the same principle as in one dimension: the densely packed state $`\rho =\overline{\rho }=1`$ inside the domain is surrounded by a $`\delta `$-function layer of strength $`\beta `$. At first glance, the asymptotic answer (43) seems not correspond to the exact answer (35) for circular, elliptic and parabolic shapes. Consider, however, the case of a circular patch as an example. Assume that the boundary of the patch being at $`r=r_0`$. According to (35), the strength of $`\delta `$-function on the boundary is $$f=\beta \frac{K_0^{}\left(r_0/\beta \right)}{K_0\left(r_0/\beta \right)}=\beta \frac{K_1\left(r_0/\beta \right)}{K_0\left(r_0/\beta \right)}$$ The ratio of Bessel functions converges rapidly to $`1`$ for $`\beta 0`$. In fact, this ratio is extremely close to $`1`$ already when $`\beta <r_0`$. Since $`\beta `$ is assumed small, for a patch of any reasonable size (larger than several units of $`\beta `$), the approximate result (43) is valid to high accuracy (exponential in $`\beta `$). In general, we can only postulate that the accuracy of (43) should be $`O(\beta ^2)`$, since we neglected the local curvature of the surface. Introduction of local curvature effects will change the asymptotic result (43), but this correction is bound to be small (of the order $`\beta ^2`$) when $`\beta 0`$. ### 6.3 Numerical simulation of states in two dimensions To illustrate these exact and asymptotic results, we have performed fully nonlinear numerical simulations for $`\alpha =1`$, $`\beta =0.1`$ and $`D=0.01`$ in two dimensions. The results of this simulations are shown in Fig. 8. gaussian radial distribution. Evolution due to (6,7) deforms this shape into a flat-top elliptical shape, which is reminiscent of the solutions derived in Sec.6.1. We conjecture that elliptical shapes are more stable than circular shapes, although a detailed analysis of the two-dimensional stability and selection mechanism has yet to be completed. ## 7 Conclusion, open problems and further applications A new model was proposed and analyzed for the collective aggregation of finite-size particles driven by the force of mutual attraction. Starting from smooth initial conditons, the solution for the particle density in this model was found to collapse into a set of delta-functions (clumps), and the evolution equations for the dynamics of these clumps were computed analytically. The energy derived for this model is well defined even when density is supported on $`\delta `$-functions. The mechanism for the formation of these $`\delta `$-function clumps is the nonlinear instability governed by the Ricatti equation (19), which causes the magnitude of any density maximum to grow without bound in finite time. At first sight, it may seem that the emergence of $`\delta `$-function peaks in the solution might be undesirable and perhaps should be avoided. However, these $`\delta `$-functions may be understood as clumps of matter, and the model guarantees that any solution eventually ends up as a set of these clumps. Subsequently, further collective motion of these clumps may be predicted using a (rather small-dimensional) system of ODEs, rather than dealing with the full non-local PDEs. The question of how many clumps arise from a given initial condition remains to be considered. One may conjecture that clump formation is extensive; so that each clump forms from the material within the range of the potential $`\mathrm{\Phi }`$, determined by $`G`$. On a longer time scale, the clumps themselves continue to aggregate, as determined by the collective dynamics (13) of weak solutions (12). This clump dynamics is also a gradient flow; so that eventually only one clump remains. A comparison with point vortex solutions of Euler’s equations for ideal hydrodynamics in two-dimensions may be made here. Prediction of the incompressible motion of an ideal fluid is governed by a set of nonlinear PDE in which pressure introduces non-locality. A drastic simplification of motion occurs, when all the vorticity is concentrated in delta-functions (point vortices) . The motion of point vortices lies on a singular invariant manifold: if started with a set of point vortices, the fluid structure will remain a set of point vortices. However, a smooth initial condition for vorticity *does not* split into point vortices under the Euler motion. In the present model, though, the physical attraction drives any initial distribution of density towards a set of delta-functions, so one is *guaranteed* to obtain effectively finite-dimensional behavior in the system after a rather short initial time. Of course, this time depends of the precise form of the long range attraction. Because two scales are present in the smoothing functions $`H`$ and $`G`$, the effects of boundary conditions warrant further study. For example, the $`H\delta `$ limit should also allow formation of boundary layers. These boundary issues were avoided in the present treatment by using periodic boundary conditions. However, it would be natural in some physical situations to apply, e.g., Neumann boundary conditions to the particle flux $`𝐉`$. The issue of boundary conditions will arise and must be addressed on an individual basis in specific applications of these equations in physics, chemistry, technology and biosciences. Our approach involving a variational principle, energy and competition of length scales is relevant for many areas of science. In particular, we have recently learned that our variational approach is applicable to bio-sciences, in particular, to the theory of insect swarming. In a recent work, Topaz et al. have discovered the formation of isolated patches of matter (*swarms*) in a variant of (6,7) with a *nonlinear* instead of *nonlocal* diffusion. The variational energy (8) for their model was $$E[\rho ]=\rho G\rho \frac{D}{2}\rho ^2$$ This case could be considered as an extension of our model for $`H(x)=\delta (x)`$, $`\mu (\overline{\rho })=1`$ and $$E[\rho ]=\rho G\rho DQ(\overline{\rho })$$ with $`Q^{}(\overline{\rho })0`$, in particular, $`Q(\overline{\rho })=\overline{\rho }^2/2`$. Alternatively, a variant of this nonlinear diffusion model can be derived if we start with $`\mu (\overline{\rho })=1\overline{\rho }`$ and $`D=0`$. We can then rewrite (6) as $$\frac{\rho }{t}+\text{div}\left(\rho \mathrm{\Phi }\right)=\text{div}\left(\rho \overline{\rho }\mathrm{\Phi }\right),$$ which, again, is a modification of swarming models considered in with non-local diffusion. We emphasize that, in our opinion, nonlocality in diffusion is advantageous, since it allows for simple generalized solutions in terms of constants and $`\delta `$-functions. These generalized solutions dominate both the dynamics and statics of the problem and greatly simplify the analytical treatment. Acknowledgments. We thank S. R. J. Brueck, P. Constantin, B. J. Geurts, J. Krug and E. S. Titi for encouraging discussions and correspondence about this work. DDH is grateful for partial support by US DOE, under contract W-7405-ENG-36 for Los Alamos National Laboratory, and Office of Science ASCAR/AMS/MICS. VP was partially supported by funding from US DOE, DE-FG02-04ER-46119 and Petroleum Research Grant 40218-AC9. Figure Captions 1. Scanning Electron Microscope (SEM) figures of self-assembly of particles in nano-channels, courtesy of S. R. J. Brueck and D. Xia. Note fully dense clumps separated by voids, evident in cases B, C and F. Cross-width of the channels as well as distance between the channels is 100 nm. 2. Numerical simulation demonstrates the emergence of particle clumps, showing formation of density peaks in a simulation of the initial value problem for equation (6) using smooth initial conditions for density with $`l=1`$, $`G(x)=H(x)=\mathrm{exp}(|x|)`$, $`\mu (\overline{\rho })=1`$. The vertical coordinate represents $`\overline{\rho }=H\rho `$, which remains finite even when the density forms $`\delta `$-functions. 3. Position of inflection point for simulation shown on Fig. 2. The dashed line in the bottom shows the minimum value allowed by the finite resolution of the mesh. In this case particular case, this minimum value is equal 0.1 with the length of the interval being 10. 4. Evolution of a Gaussian initial condition for $`\rho `$(x,0) with $`\mu (\overline{\rho })=1\overline{\rho }`$ and $`H=\mathrm{exp}(|x|/ϵ)/(2ϵ)`$ where $`ϵ=\alpha /10`$. The solution quickly forms a plateau of maximal possible density ($`\rho _{\text{max}}=1`$). 5. Stationary equilibrium solution for $`\alpha =1`$, $`\beta =0.1`$ and $`L=1`$. Top $`\rho (x)`$ (with representation of $`\delta `$-functions at $`x=\pm L`$ as vertical red lines). Middle: $`\mathrm{\Phi }=(G\rho )(x)`$ for the same solution. Bottom: $`\overline{\rho }=H\rho `$ for the same solution. 6. Stationary jammed states for $`\alpha =1`$, $`\beta =0.1`$ and $`L=1`$. Top $`\rho (x)`$ (again, $`\delta `$-functions at $`x=\pm L`$ are represented by vertical red lines). Middle: the potential $`\mathrm{\Phi }=(G\rho )(x)`$. Bottom: $`\overline{\rho }=H\rho `$. 7. Convergence to analytic equilibrium solution (circles) for different times (colored solid lines, see legend). In simulation, $`D=0.01`$, $`\alpha =1`$, $`\beta =0.1`$ 8. Two-dimensional evolution of a gaussian initial profile for consecutive times with $`D=0.01`$, $`\alpha =1`$, $`\beta =0.1`$
warning/0506/quant-ph0506219.html
ar5iv
text
# An Introduction to Quantum Game Theory ## .1 Some background history Game theory traditionally began in 1944 with *The Theory of Games and Economic Behavior*, by *John von Neumann* and *Oscar Morgenstern*. But it had antecedents stemming from the Hungarian mathematician von Neumann’s earlier simultaneous interest in game theory and the foundations of quantum mechanics. Since we are interested in quantum games, we will describe the development briefly as follows. In 1900 *Max Planck*, attempting to get rid of the infinite energy implied in the then current formula for black body radiation, proposed a solution in which electromagnetic radiation energy was only emitted or absorbed in discrete energy units or *quanta*, multiples of a fundamental unit $`h`$: $`h\nu ,2h\nu ,3h\nu \mathrm{}`$, where $`\nu `$ is the frequency of the radiating oscillator, and $`h`$ is now known as Planck’s constant. In 1905 *Albert Einstein* used Planck’s quantum as an explanation for the photoelectric effect, whereby metals required incident light of a minimum frequency before they would release electrons. Incident light of frequency $`\nu `$ appeared to behave as a collection of particles (‘photons’), each with energy $`E=h\nu `$. *Niels Bohr* then developed a useful, if unsatisfactory, model of the atom as a nucleus surrounded by planetary electrons whose orbits assumed only discrete values for the angular momentum, corresponding to multiples of Planck’s quantum of energy: $`\frac{h}{2\pi },\frac{2h}{2\pi },\frac{3h}{2\pi },\mathrm{}`$. In 1924 *Louis de Broglie* helped clarify the picture by associating with matter a wave, and noting that waves in closed loops, such as the electron ‘circling’ the nucleus, were required to fit evenly around the loop—i.e. to have *whole number* cycles. The whole numbers $`1,2,3,\mathrm{}`$ were thus associated with Planck’s quanta (times a constant $`a`$): $`1ah,2ah,3ah,\mathrm{}`$. This was the *old* quantum theory. The *new* quantum theory began in 1925 when *Werner Heisenberg* conceived of representing physcial quantities by sets of time-dependent *complex* numbers. Heisenberg’s *matrix mechanics* essentially involved $`N\times N`$ input-output matrices $`H`$, representing transitions between states of matter. If we denote by $`\psi `$ the state of the system we are interested in at time $`t`$(we will for the moment set $`t`$ to zero), where $`\psi `$ is a $`N\times 1`$ vector, then Heisenberg was working with the eigenvector-eigenvalue system $$H\psi =E\psi $$ (1) where $`E`$, a scalar, represents some quantized energy level. Assuming the system of $`N`$ equations is nondegenerate, there are $`N`$ solutions for $`E`$, say $`E_n,\text{ }n=1,2,\mathrm{},N`$. The $`E_n`$ eigenvalues, or energy levels, are associated with an $`N`$-eigenvector-basis for the state space of $`\psi `$. The following year *Erwin Schrödinger*, looking for an electromagnetic interpretation of the same phenomena, published his famous wave equation $$i\mathrm{}\frac{\psi }{t}=\frac{\mathrm{}^2}{2m}\left(\frac{^2}{x^2}+\frac{^2}{y^2}+\frac{^2}{z^2}\right)\psi +V\psi ,$$ (2) where $`i=\sqrt{}1`$, $`\mathrm{}`$ is Planck’s quantum of energy $`h`$ divided by $`2\pi `$, and $`V`$ is potential energy. To Schrödinger’s delight, he discovered that his approach and Heisenberg’s matrix mechanics were mathematically equivalent, one form of this equivalence being suggested by the equation $`i\mathrm{}\frac{\psi }{t}=H\psi `$. If we, for example, set $`\psi =Aexp^{(i\frac{E}{\mathrm{}}t)}`$ in Schrödinger’s equation (2), and let $`H=\frac{\mathrm{}^2}{2m}\left(\frac{^2}{x^2}+\frac{^2}{y^2}+\frac{^2}{z^2}\right)+V`$, then we obtain $`E\psi =H\psi `$, which is Heisenberg’s equation (1). A few years later *John von Neumann*, whose interest in quantum mechanics was inspired by Heisenberg, ‘showed that quantum mechanics can be formalized as a calculus of Hermitian operators in Hilbert space and that the theories of Heisenberg and Schrödinger are merely particular representations of this calculus.’ (Jammer, , p.22) Recall that a *Hermitian matrix* is one that is its own complex-conjugate transpose. For example, consider the matrix $`\sigma _y=\left(\begin{array}{cc}0& i\\ i& 0\end{array}\right)`$. The transpose of this matrix is $`\sigma _y^T=\left(\begin{array}{cc}0& i\\ i& 0\end{array}\right)`$. Then if we take the complex conjugate, by changing the signs of the imaginary parts, $`ii,ii`$, we again obtain the matrix $`\sigma _y`$. So $`\sigma _y`$ is Hermitian. A Hermitian matrix may be considered an *operator* on a vector in Hilbert space. Recall that Hilbert space is simply a vector space defined over the complex numbers $`𝐂`$, with a defined *norm* or *length* or *inner product*. For the vector $`\psi `$ the norm is $`\psi =\sqrt{\psi ^{}\psi }`$, where $`\psi ^{}`$ is the complex conjugate transpose of $`\psi `$. Hilbert spaces may be infinite dimensional, but we will only consider finite dimensional spaces in this essay. It was during this heady period that game theory arose. The name ‘game’ was introduced in 1921 by the French mathematician Emil Borel, who was preoccupied with bluffing in poker and initiated ‘la théorie du jeu’. In his 1928 paper vN28 , written for Karl Menger’s Vienna Colloquium, von Neumann defined, and completely solved, two-person zero-sum games. He speculated on $`N`$-person games, which were more complicated due to the possibility of coalitions: with three people or more, some people could benefit from cooperation. Later, in a famous paper delivered to the Princeton economics club in 1932, the same year his book on the foundations of quantum mechanics was published, von Neuman laid out the whole apparatus of linear programming and the foundations of his later game theory book with Morgenstern. (This paper was not published util 1937 vN37 .) Central to many results was the linear programming problem and its dual DGale . The linear programming problem is this: given an $`m\times n`$ matrix $`A`$, an $`n\times 1`$-vector $`b`$, and an $`m\times 1`$-vector $`c`$, find a non-negative $`m\times 1`$-vector $`x`$ such that $$x^Tc\text{ is a maximum }$$ (3) subject to $$x^TAb^T.$$ (4) The dual problem is that of finding a non-negative $`n\times 1`$-vector $`y`$ such that $$y^Tb\text{ is a minimum }$$ (5) subject to $$Ayc.$$ (6) The only major game theoretic result missing from von Neumann-Morgenstern (and indeed one missing from the quantum game theory literature) is the theory of the core (LR, , chapter 8). The *core* arises in $`N`$-person game theory. In $`N`$-person game theory players’ interests are not necessarily opposed, Some players may improve their (expected) payoffs by forming coalitions with other players. A maximum value can be determined for each subset of players, which gives rise to the *characteristic function* of the game. Let $`S`$ be a member of the set of subsets of $`N`$. The characteristic function $`v(S)`$ is a mapping from the set of subsets (i.e. coalitions) of players to an (expected) payoff value in the set of real numbers $`R`$: $$v(S):SR.$$ (7) The value $`v(S)`$ is determined as the maximum value obtainable by $`S`$ in the two-person game between the coalition $`S`$ and the coalition of all remaining players $`NS`$. An *imputation* is a set of numbers (allocations or payoffs) $`\{\pi _i\}`$ assigned to each player $`i`$ in $`N`$. The core $`C_x`$ is the set of imputations $`C_x=\{\{\pi _i\}_x\}`$ such that $$v(S)\underset{iS}{}\pi _i\text{ for every subset }S\text{ in }N,\text{ and }\underset{iN}{}\pi _i=v(N).$$ (8) The core (it may be empty) is critical to economic equilibrium. The core restricts the value of any coalition to be not greater than the sum of the imputed payoffs to each member of the coalition individually. Debreu and Scarf DeSc showed that in a replicated market game the core shrinks down to a set of imputations which can be interpreted in terms of a price system emerging as its limit. Meanwhile, in quantum mechanics, the reactionary forces of determinism were at work. In a 1935 paper EPR *Einstein-Podolsky-Rosen* (EPR) attempted to prove the *incompleteness* of quantum mechanics by considering entangled pairs of particles which go off in different directions. The particles may become separated by light-years. Nevertheless a measurement of one particle will instantly affect the state of the other particle, an example of quantum mechanics’ ‘spooky action at a distance’. (We will discuss entanglement later, in the body of this essay, but essentially two particles are entangled if their wave functions cannot be written as tensor products.) This instantaneous effect is sometimes called the ‘EPR channel’, though properly speaking it should be called the *Bohr channel* because Bohr argued for its existence, while EPR argued against it. *John Bell* JSB formulated a set of inequalities that would distinguish experimentally whether quantum mechanics was incomplete, or whether physics is *non-local*, permitting instantaneous propagation of some effects of some causes. Fortunately Bohr was right and EPR were wrong, as experimental evidence has decisively demonstrated.NG2005 The Bohr channel is now the basis of quantum teleportation, and, indeed, every quantum computer is in some sense a demonstration of the Bohr effect. As it stands today, quantum game theory can probably be viewed as a subbranch of quantum computation. With respect to the latter development, it was apparently *Richard Feynman* RF who first foresaw the unusual power of quantum computers, noting that simulation of quantum evolution in a classical computer would invole an exponential slowdown in time. Once again there is a direct line from von Neumann vN56 (with Stan Ulam SU ): ‘In the nineteen fifties, Ulam and von Neumann began to discuss computational models known as cellular automata, in which simple rules of computation applied to systems with many degrees of freedom could produce complex patterns of behavior. By the nineteen eighties, Friedkin, Feynman, Minsky and others were speculating on the possibility of describing the laws of physics and the universe in terms of cellular automata and computation. Underlying their ideas was a dissatisfaction with the conventional description of physics based on continuous space and time.’ JR *David Deutsch* DD suggested that quantum *superposition* might allow the parallel performance of many classical computations. Indeed, we shall see that superposition is the key new ingredient that makes quantum games different from classical games, whether or not the superposed states are *entangled*. For dynamic games, superposition suffices, though static games generally require entanglement also. (Superposition is the ability of a quantum observable to be in a linear combination of two or more states at the same time.) The ‘killer app’ that created a storm of interest in quantum computation came when *Peter Shor* PS showed that a quantum mechanical algorithm could factor numbers in polynomial time. This was an exponential speed-up over factoring algorithms available to classical computers. Shor’s algorithm relies mainly on superposition and an ingenious application of the quantum Fourier transform. Another result was obtained by *Lov Grover* LG , who showed a quantum mechanical way to speed up the search for items in an $`N`$-item database from $`O(N)`$ steps to $`O(\sqrt{N})`$ steps. Grover’s result is based upon the rotation of quantum states (vectors) in Hilbert space. Quantum game theory seems to have crystallized when *David Meyer* gave a talk on the subject at Microsoft Corporation (see DM for an account). Of the twelve quantum games considered in this essay, three are due to Meyer (the Spin Flip game, and Guess a Number games I and II). As von Neumann and Morgenstern noted vNM , ‘In order to elucidate the conceptions which we are applying to economics, we have given and may give again some illustrations from physics. There are many social scientists who object to the drawing of such parallels on various grounds, among which is generally found the assertion that economic theory cannot be modeled after physics since it is a science of social, of human phenomena, has to take psychology into account, etc. Such statements are at least premature.’ One may conversely note that some may similarly object to mixing economic concepts with those of quantum mechanics, but such objections are at least premature. Indeed, the human brain is arguably a quantum computer HS1 HS2 RP DD89 DD2002 , though the mind may be more than that, so to ignore quantum mechanics in questions of psychology, much less economics, is folly indeed. In the reverse direction, the role of the human mind in the quantum measurement problem has been a subject of contention JJ since it was first clearly delineated by von Neumann. In any event, quantum games may have lessons both for economics and quantum mechanics. ## .2 Preliminary mathematical pieces Before defining a game, we are going to give an example of one. This example, the Spin Flip Game in the next section, will highlight some of the differences between traditional game theory and quantum game theory. In order to explain how the Spin Flip Game works, we will need some modest mathematical preliminaries, involving $`2\times 1`$ vectors and $`2\times 2`$ matrices. The following simple vectors will prove quite useful for our purposes: $$u=\left(\begin{array}{c}1\\ 0\end{array}\right),d=\left(\begin{array}{c}0\\ 1\end{array}\right).$$ (9) These are, of course, *basis* vectors for 2-dimensional (complex) space, as any point can be expressed in the form of $`au+bd`$ (where, in general, it is assumed that $`a`$ and $`b`$ are complex scalars, $`a,b𝐂`$). But $`u`$ and $`d`$ can also represent many ’spaces’ or states outside geometry: Yes or No responses, Up or Down spin states of an electron (with spin measured in the $`z`$ direction), Heads or Tails in a probability sequence, Success or Failure of a bidding process or an electronic device, and so on. A choice of $`u`$ or $`d`$ can also represent player moves in a game, and we can represent a sequence of such moves by the *bits* in a binary number, or the quantum equivalent *qubits*. Bits and qubits differ by the fact that a bit $`b`$ is a single number, $`b\{0,1\}`$, while a qubit $`q`$ is a vector in a two-dimensional Hilbert space, $`q\{au+bd\}`$. (Later we will introduce the Dirac notation $`|0`$, $`|1`$, and in this essay there is the correspondence $`u`$ $`|u\left(\begin{array}{c}1\\ 0\end{array}\right)|0\text{ bit }0`$, and the similar correspondence $`d`$ $`|d\left(\begin{array}{c}0\\ 1\end{array}\right)|1\text{ bit }1`$. For example, to foreshadow what is to come, the 5-qubit register or sequence $`|10011`$ could represent the tensor product of vectors as well as the number $`19\text{ }(=2^4+2^1+2^0)`$: $$|10011=\left(\begin{array}{c}0\\ 1\end{array}\right)\left(\begin{array}{c}1\\ 0\end{array}\right)\left(\begin{array}{c}1\\ 0\end{array}\right)\left(\begin{array}{c}0\\ 1\end{array}\right)\left(\begin{array}{c}0\\ 1\end{array}\right)$$ (10) $`=(0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,0,1,0,0,0,0,0,0,0,0,0,0,0,0)^T`$. In the latter vector, the $`1`$ is in the 20th slot, not the 19th, because we start counting from $`0`$, which occupies the first slot. The same sequence could have also been written $`duudd`$.) Next we need some way to transform one state into another. For a two-state system, it is useful to do this with the Pauli spin matrices. The three $`2\times 2`$ *Pauli spin matrices* are $$\sigma _x=\left(\begin{array}{cc}0& 1\\ 1& 0\end{array}\right),\sigma _y=\left(\begin{array}{cc}0& i\\ i& 0\end{array}\right),\sigma _z=\left(\begin{array}{cc}1& 0\\ 0& 1\end{array}\right).$$ (11) These three matrices, along with the following unit matrix $`\mathrm{𝟏}`$, $$\mathrm{𝟏}=\left(\begin{array}{cc}1& 0\\ 0& 1\end{array}\right),$$ (12) span $`2\times 2`$ Hermitian matrix space (recall that a Hermitian matrix has diagonal elements that are real, and mirror-image off-diagonal elements that are complex conjugates of each other). Each of the spin matrices has a simple effect on the base states $`u`$ and $`d`$. In particular, $$\mathrm{𝟏}u=u,\text{ }\mathrm{𝟏}d=d$$ (13) $$\sigma _xu=d,\text{ }\sigma _xd=u$$ (14) $$\sigma _zu=u,\text{ }\sigma _zd=d.$$ (15) Table 1 summarizes some matrix properties of the Pauli spin matrices: ## .3 The spin flip game Electrons have two spin states: spin up and spin down. Let us consider a simple game of electron spin flip played between Alice and Bob. Alice first prepares the electron in spin up state $`u`$. After this initial step, Bob applies either the $`\sigma _x`$ or the $`\mathrm{𝟏}`$ matrix to $`u`$, resulting in either $$\sigma _xu=d\text{ or }\mathrm{𝟏}u=u.$$ (16) Then Alice (not knowing Bob’s action or the state of the electron) takes a turn, also applying either $`\sigma _x`$ or $`\mathrm{𝟏}`$ to the electron spin. Then Bob (not knowing Alice’s action or the state of the electron) takes another turn. Finally, the electron spin state is measured. If it is in the $`u`$ state, Bob wins $1, and Alice loses $1. If it is in the $`d`$ state, Alice wins $1, while Bob loses the same amount. The sequence of possible choices by Bob (*columns*) and Alice (*rows*) are summarized in Table II. Note that Alice’s move is the middle one in each sequence of three, reading from right to left. For example $`\mathrm{𝟏},\mathrm{𝟏},\sigma _x`$ means that Bob played $`\sigma _x`$, followed by Alice’s play of $`\mathrm{𝟏}`$, followed by Bob’s play of $`\mathrm{𝟏}`$. The net result is $`\mathrm{𝟏𝟏}\sigma _xu=d`$. Thus Alice wins $1. The sequence of spin states after each move, starting from the initial $`u`$ state are shown in Table III. Again, each sequence of three should be read from right to left. Finally, Table IV shows the payoff to *Alice*, positive if the final spin is in the $`d`$ state, negative if it is in the $`u`$ state. This is the basic Spin Flip Game, which we are going to extend in two directions: first, by considering probabilistic moves, and, second, by considering *quantum superposition* (without *quantum entanglement*) of states. But before doing this, let’s consider some basic game theory terminology. ## .4 First game definitions and strategies As is implicit in the previous section, a *game* $`\mathrm{\Gamma }`$ may be defined as a set $`\mathrm{\Gamma }=\mathrm{\Gamma }`$(players,moves or actions,outcomes,payoffs). In the Spin Flip Game, the *players* were Alice and Bob, the *moves* were the application of the matrices $`\sigma _x`$ or $`\mathrm{𝟏}`$, the *outcomes* were the spin states $`u`$ or $`d`$, and the *payoffs* to Alice were either +1 or -1, according to whether the final state was $`d`$ or $`u`$, respectively. Since this was a *two-person, zero-sum* game, the payoffs to Bob were the exact opposite of those to Alice. Omitted thus far in the account of the game is any explanation how Alice and Bob determined their moves—how they decided whether to play $`\sigma _x`$ or $`\mathrm{𝟏}`$. A *strategy* is a rule for determining a move at any stage of a game. That is, in our example, a *move* is a member of the set $`\{\mathrm{𝟏},\sigma _x\}`$, while a *strategy* is a *function* $`f`$ mapping the state of the game to the set of moves: $`f:\text{ game state }\{\mathrm{𝟏},\sigma _x\}`$. (There seems to be confusion on this point in the quantum game theory literature.) This is not quite a good definition, since the ‘state of the game’ may not be known to a player; a player may know little more than his or her move. So let’s revise this to: a *strategy* for Alice is a mapping $`f_A:\{\text{ Alice’s information }\}\{\text{ Alice’s moves }\}`$. Similarly for Bob. In the Spin Flip Game Alice, after initial preparation of the electron, has only one opportunity to choose a move, so she has a single strategy at the second, or middle, step of the sequence of three moves. Bob has strategies for the first and last steps. Thus, associated with a sequence of moves is a sequence of strategies. In economics, strategies are highly dependent on a player’s *information*. Of particular interest is *asymmetric* information, where one player has some information advantage over another, or where the information sets of the players are not the same. If Bob can make quantum moves that Alice cannot, then clearly Bob has an information advantage in at least that respect. Strategies are endogenous to a game, given the game’s allowed moves and payoffs, so strategies are not properly part of the game’s definition. Rather, solving a game essentially means determining the optimal strategies for the players. The concept of information set is important. In the Spin Flip Game we said that neither Bob nor Alice could know the other person’s moves. Suppose we relaxed this assumption. Then Alice would know Bob’s first move, and could choose her move accordingly, but it would make no difference. Bob, seeing Alice’s move (and knowing his own first move), could always choose a final move that would leave the electron in a spin up state $`u`$. He would win 100 percent of the time. It would not be a ‘game’, but rather a racket. So in this case we must limit the information sets of Alice and Bob in order to make it a game in the first place. Now, as an example let us consider the following strategies, $`f_A`$ and $`f_B`$, for Alice and Bob, respectively. These will be called *mixed* strategies because they involve selection of a move with some probability mechanism. $`f_A`$ $`=`$ $`\text{play }\mathrm{𝟏}\text{ with probability }p={\displaystyle \frac{1}{2}},\text{ play }\sigma _x\text{ with probability }q={\displaystyle \frac{1}{2}}`$ (17) $`f_B`$ $`=`$ $`\text{play }\mathrm{𝟏}\text{ with probability }p={\displaystyle \frac{1}{2}},\text{ play }\sigma _x\text{ with probability }q={\displaystyle \frac{1}{2}}.`$ (18) Then, looking at the columns of Table IV, we see that Alice’s *expected payoff* $`\overline{\pi }_A`$, no matter what Bob does, is always $$\overline{\pi }_A=\frac{1}{2}(+1)+\frac{1}{2}(1)=0$$ (19) while, looking at the rows of Table IV, Bob’s expected payoff is always $$\overline{\pi }_B=\frac{1}{4}(+1)+\frac{1}{4}(1)+\frac{1}{4}(1)+\frac{1}{4}(+1)=0.$$ (20) Of course, for the concept of *mixed* strategies and *expected* payoffs to make much sense, we should consider a sequence of $`N`$ games $$\mathrm{\Gamma }_N\mathrm{\Gamma }_{N1}\mathrm{\Gamma }_{N2}\mathrm{}\mathrm{\Gamma }_3\mathrm{\Gamma }_2\mathrm{\Gamma }_1.$$ (21) The *actual* payoff to Alice, letting $`x`$ stand for the number of wins in $`N`$ games, will be a member of the *payoff set* $$\mathrm{\Pi }=\{f(x;N)\}=\{2xN,\text{ for }x=0,1,\mathrm{},N\}$$ (22) while the probability of these payoffs are $$P(\mathrm{\Pi })=\{f(x;N,p)\}=\{\left(\begin{array}{c}N\\ x\end{array}\right)p^xq^{Nx},\text{ for }x=0,1,\mathrm{},N\}.$$ (23) For example, with $`N=3`$, the possible payoffs to Alice are $`\{3,1,1,3\}`$, and if $`p=\frac{1}{2}`$ these have respective probabilities $`\{\frac{1}{8},\frac{3}{8},\frac{3}{8},\frac{1}{8}\}`$. Alice’s expected payoff $`\overline{\pi }_A`$ is $`0`$, but if $`N`$ is odd, her actual payoff will never be $`0`$. Physicists will recognize equation (22) as giving the possible outcome states when a massive particle of spin $`\frac{N\mathrm{}}{2}`$ is measured. The spin in this case defines an $`(N+1)`$-state quantum system, with possible outcomes for the spin values (in terms of the fundamental unit $`\frac{\mathrm{}}{2})`$ given by equation (22). Thus the *measured* spin states of the massive particle may be thought of as being determined by $`N`$ Spin Flip games between Alice and Bob. In the matrix of payoffs analogous to Table IV, for a general two-person, zero-sum game, let Alice’s moves be represented by the mixed strategy (the set of probabilities over moves) $`P_A=\{a_1,a_2,\mathrm{},a_m\}`$, while the mixed strategy of Bob is represented by $`P_B=\{b_1,b_2,\mathrm{},b_n\}`$. Let the payoffs to Alice be represented by the $`m\times n`$ matrix $`[\pi _{ij}]`$. Then the *expected payoff* to Alice is $$\overline{\pi }_A=\underset{j=1}{\overset{n}{}}\underset{i=1}{\overset{m}{}}\pi _{ij}a_ib_j.$$ (24) In this context, we should mention the *minimax theorem* which says that for every finite two-person, zero-sum game $$max_{_{P_A}}(min_{_{P_B}}\overline{\pi }_A)=min_{_{P_B}}(max_{_{P_A}}\overline{\pi }_A).$$ (25) That is, Alice chooses probable moves to maximize her expected payoff, while Bob choses probable moves to minimize Alice’s expected payoff. The minimax theorem says the payoff to Alice’s maximizing set of probabilities given Bob’s minimizing set of probabilites is equal to the payoff to Bob’s minimizing set of probabilities given Alice’s maximizing set of probabilities. ## .5 Amplitudes and superpositions and his cheatin’ heart Let’s consider a quantum state (a vector) $`\psi `$ of the following form, where $`a`$ and $`b`$ may be complex scalars: $$\psi =au+bd$$ (26) In quantum computation, this superimposed two-dimensional state is known as a *qubit*, which we will discuss in detail later. Here $`a`$ and $`b`$ are *amplitudes*, and a (von Neumann) measurement of $`\psi `$ will obtain the base state $`u`$ with probability $`|a|^2`$, while the measurement will yield base state $`d`$ with probability $`|b|^2`$, where $`|a|^2+|b|^2=1`$. (Recall that for a complex number $`a`$, and its complex conjugate $`a^{}`$, we have $`aa^{}=a^{}a=|a|^2`$.) This raises the possibility of games, including variants of the Spin Flip Game, for which there is no classical analog. For example, set $`a=b=\frac{1}{\sqrt{2}}`$. Then the probability of either $`u`$ or $`d`$ is $`|\frac{1}{\sqrt{2}}|^2=\frac{1}{2}`$. Thus probability is built into measurements of the state vector, irrespective of whether a mixed strategy is chosen by either Bob or Alice. Here $`u`$ and $`d`$ are orthonormal (that is, the inner product of $`u`$ with $`d`$ is 0, and the inner product of either $`u`$ or $`d`$ with itself is 1), so we may obtain $`a`$ as the inner product $$\psi ,u=au,u+bd,u=a(1)+b(0)=a.$$ (27) A similar computation will yield $`b`$. $`\mathrm{𝐀𝐥𝐢𝐜𝐞}\text{ }\mathrm{𝐂𝐡𝐞𝐚𝐭𝐬}.`$ Now let us consider a variation of the Spin Glip Game—let’s call it *Alice Cheats*—in which Alice has a way of cheating in the initial preparation of the spin state of the electron. First, suppose she initially prepares the electron in spin state $`d`$, knowing that Bob thinks it will be in spin state $`u`$. Otherwise the game is exactly as before: both Bob and Alice play either $`\mathrm{𝟏}`$ or $`\sigma _x`$. It is easy to see that the arrangement of spin states changes in Table III, and the arrangemnt of payoffs to Alice changes in Table IV, but the set of payoffs $`\mathrm{\Pi }`$ is still the same, and the corresponding payoff probabilities $`P(\mathrm{\Pi })`$ to Alice are unchanged. Thus Alice has cheated to no avail. She simply changed the initial state from $`u`$ to $`d`$, and it had no impact on the outcome of the game. Where she previously got +1, she now gets -1, and vice-versa. So Alice tries something else. She choses the initial state to be $`\frac{1}{\sqrt{2}}(u+d)`$. Then whether Bob plays $`\mathrm{𝟏}`$ or $`\sigma _x`$, his move leaves the state of the game unchanged: $$\mathrm{𝟏}[\frac{1}{\sqrt{2}}(u+d)]=\frac{1}{\sqrt{2}}(\mathrm{𝟏}u+\mathrm{𝟏}d)=\frac{1}{\sqrt{2}}(u+d),$$ (28) $$\sigma _x[\frac{1}{\sqrt{2}}(u+d)]=\frac{1}{\sqrt{2}}(\sigma _xu+\sigma _xd)=\frac{1}{\sqrt{2}}(d+u).$$ (29) Since $`u+d=d+u`$, the state is unchanged by the play of either $`\mathrm{𝟏}`$ or $`\sigma _x`$. However, when the final measurement of the (unchanged) state of the electron is taken, Alice discovers to her frustration that she once more wins or loses a dollar with equal probability, because a measurement of the final superposed state yields $`u`$ or $`d`$ with equal probability. For a single game, the payoff set $`\mathrm{\Pi }`$ and corresponding probabilities $`P(\mathrm{\Pi })`$ are: $$\mathrm{\Pi }=\{1,+1\}$$ (30) $$P(\mathrm{\Pi })=\{(\frac{1}{\sqrt{2}})^2,(\frac{1}{\sqrt{2}})^2\}=\{\frac{1}{2},\frac{1}{2}\}.$$ (31) $`\mathrm{𝐁𝐨𝐛}\text{ }\mathrm{𝐂𝐡𝐞𝐚𝐭𝐬}`$. Let’s return to our basic Spin Flip Game, where a repentent Alice prepares the electron in an initial $`u`$ state, with the added detail that she follows a mixed strategy, and choses $`\mathrm{𝟏}`$ or $`\sigma _x`$ each with probability $`p=\frac{1}{2}`$. But now we allow Bob to cheat. Since Bob does not prepare the initial electron state, Bob’s method of cheating will differ from Alice’s. What dastardly things can Bob do? Bob has some extra Pauli spin matrices up his sleeve, namely $`\sigma _y`$ and $`\sigma _z`$, as well as linear combinations of these. In addition, Bob has the final move. Let’s suppose that Bob plays the so-called Hadamard operator $`H=\frac{1}{\sqrt{2}}(\sigma _x+\sigma _z)`$: $$H=\frac{1}{\sqrt{2}}\left(\begin{array}{cc}1& 1\\ 1& 1\end{array}\right).$$ (32) After Bob’s first move, the spin state would be $$Hu=\frac{1}{\sqrt{2}}\left(\begin{array}{cc}1& 1\\ 1& 1\end{array}\right)\left(\begin{array}{c}1\\ 0\end{array}\right)=\frac{1}{\sqrt{2}}\left(\begin{array}{c}1\\ 1\end{array}\right)=\frac{1}{\sqrt{2}}(u+d).$$ (33) As we saw in equations (28-29), Alice’s mixed strategy will not change this state. Then Bob plays $`H`$ again to obtain: $$H(Hu)=\frac{1}{\sqrt{2}}\left(\begin{array}{cc}1& 1\\ 1& 1\end{array}\right)\frac{1}{\sqrt{2}}\left(\begin{array}{c}1\\ 1\end{array}\right)=\frac{1}{2}\left(\begin{array}{c}2\\ 0\end{array}\right)=u.$$ (34) Bob will always win. This results from Bob’s ability to create a *superposition* of states (and his having the final move). Like Schrödinger’s cat that is simultaneously both alive and dead, the electron spin is simultaneously both $`u`$ and $`d`$ after Bob applies the Hadamard matrix $`H`$ to $`u`$. Alice cannot alter the outcome by playing a classical mixed strategy that choses a play of $`\mathrm{𝟏}`$ with probability $`p`$ and $`\sigma _x`$ with probability $`1p`$. ## .6 Guess a number games To understand the *Guess a Number Game*, we will first need to introduce some more concepts, including *qubits*, the *Walsh-Hadamard transformation* (the $`n`$-bit analogue of the Hadamard transformation) and some elements of the *Grover search algorithm* LG . The Grover search algorithm is one of the fundamental techniques of quantum computation, so it is not surprising it shows up in quantum game theory. $`\mathrm{𝐃𝐢𝐫𝐚𝐜}\text{ }\mathrm{𝐧𝐨𝐭𝐚𝐭𝐢𝐨𝐧}.\text{ }`$ For convenience, we are going to alter our designations for $`u`$ and $`d`$ into forms that will denote each $`2\times 1`$ vector and also its $`1\times 2`$ *complex conjugate* transpose: $$|u=\left(\begin{array}{c}1\\ 0\end{array}\right),u|=(1,0),|d=\left(\begin{array}{c}0\\ 1\end{array}\right),d|=(0,1).$$ (35) Note that if $`|x=\left(\begin{array}{c}1\\ i\end{array}\right),\text{ then }x|=(1,i)`$. This is the *Dirac bracket notation*, where $`x|`$ is the *bra* and $`|x`$ is the *ket*. The bras are horizontal, and the kets are vertical. Notice that we may then use the form $`|ud|`$: $$|ud|=\left(\begin{array}{c}1\\ 0\end{array}\right)(0,1)=\left(\begin{array}{cc}0& 1\\ 0& 0\end{array}\right)$$ (36) where $`|ud|`$ turns a $`|d`$ into an $`|u`$; namely, $`|ud|d=|u`$; and an $`|u`$ into a $`2\times 1`$ zero vector, namely $`|ud|u=\left(\begin{array}{c}0\\ 0\end{array}\right)`$. $`\mathrm{𝐐𝐮𝐛𝐢𝐭𝐬}.`$ Consider an $`n`$-bit binary number $`x`$: $$x=b_{n1}b_{n2}\mathrm{}b_2b_1b_0,$$ (37) where each $`b_i`$ is either $`0`$ or $`1`$, $`b_i\{0,1\}`$. Note that the decimal equivalent of $`x`$ is $$x=b_{n1}2^{n1}+b_{n2}2^{n2}+\mathrm{}+b_22^2+b_12^1+b_02^0.$$ (38) In a quantum computer, each $`b_i`$ may be represented by $`|u`$ or $`|d`$, respectively. We make the correspondence $`|u|0,|d|1`$, and call $`\{|0,|1\}`$ the *computational basis*. The latter representation, however, makes them quantum bits or *qubits*—vectors in a two-dimensional Hilbert space. Each qubit can be any linear combination $`a|0+c|1`$, where $`|a|^2+|c|^2=1`$. For example, consider the 3-qubit state $`|\psi =|q_2|q_1|q_0\text{ where}`$ (39) $`|q_2={\displaystyle \frac{1}{\sqrt{2}}}(|0+|1)`$ (40) $`|q_1=|1`$ (41) $`|q_0=|1.`$ (42) Then the quantum register is the superposition of $`|3`$ and $`|7`$: $`|\psi ={\displaystyle \frac{1}{\sqrt{2}}}(|0+|1)|1|1`$ (43) $`={\displaystyle \frac{1}{\sqrt{2}}}(|011+|111)`$ (44) $`={\displaystyle \frac{1}{\sqrt{2}}}(|3+|7).`$ (45) This calculation will be further clarified below. A collection of $`n`$ qubits is called a *quantum register* of size $`n`$. There are $`N=2^n`$ such numbers or quantum register states $`x`$ in terms of the computational basis $`b_i`$, $`b_i\{|0,|1\}`$; hence $`xS=\{0,1,2,\mathrm{},N1\}`$. So our Hilbert space has dimension $`N=2^n`$. That is, a classical computer with $`n`$ bits has a total of $`2^n`$ possible states. By contrast, a quantum computer with $`n`$ qubits can be in any superposition of these $`2^n`$ states, which results in an arbitrary state or vector in $`2^n`$-dimensional Hilbert space. A superposition $`|\psi _s`$ of *all* the computational basis states, letting $`a_x`$ be the probability amplitude associated with the number or state $`x`$, would be designated $$|\psi _s=\underset{x=0}{\overset{2^n1}{}}a_x|x.$$ (46) If all amplitudes $`a_x`$ are equal, then this superposition is designated $$|\psi _s=\frac{1}{\sqrt{2^n}}\underset{x=0}{\overset{2^n1}{}}|x.$$ (47) Note that in the summation in equation (47), $`|x`$ runs through all basis states or numbers, and all the basis states are orthogonal to each other. Hence for a given number or state $`|z`$, we have that the amplitude for $`|z`$ is the inner product $$z|\psi _s=\frac{1}{\sqrt{2^n}}.$$ (48) A measurement of $`|\psi _s`$ will thus yield $`|z`$ with probability $$|z|\psi _s|^2=\frac{1}{2^n}.$$ (49) Now, when we have a *many-state* system of $`|u`$s and $`|d`$s (i.e., $`|0`$s and $`|1`$s) like this, each in a Hilbert space $`𝐇_2`$ of $`2`$ dimensions, we simply place the states side by side. Two such states side by side form a Hilbert space of $`𝐇_4=𝐇_2𝐇_2`$ dimensions. Basis vectors in a $`2`$-qubit quantum register could thus be represented $$|0|0=|u|u=\left(\begin{array}{c}1\\ 0\end{array}\right)|u=\left(\begin{array}{c}u\\ \mathrm{𝟎}\end{array}\right)=\left(\begin{array}{c}1\\ 0\\ 0\\ 0\end{array}\right).$$ (50) $$|0|1=|u|d=\left(\begin{array}{c}1\\ 0\end{array}\right)|d=\left(\begin{array}{c}d\\ \mathrm{𝟎}\end{array}\right)=\left(\begin{array}{c}0\\ 1\\ 0\\ 0\end{array}\right).$$ (51) $$|1|0=|d|u=\left(\begin{array}{c}0\\ 1\end{array}\right)|u=\left(\begin{array}{c}\mathrm{𝟎}\\ u\end{array}\right)=\left(\begin{array}{c}0\\ 0\\ 1\\ 0\end{array}\right).$$ (52) $$|1|1=|d|d=\left(\begin{array}{c}0\\ 1\end{array}\right)|d=\left(\begin{array}{c}\mathrm{𝟎}\\ d\end{array}\right)=\left(\begin{array}{c}0\\ 0\\ 0\\ 1\end{array}\right).$$ (53) Physicists, who get bored with the excessive notation, usually compress the tensor product of qubits as $$|u|u\mathrm{}|u|u|u\mathrm{}|u.$$ (54) And then often compress it again: $$|u|u\mathrm{}|u|uu\mathrm{}u.$$ (55) All these different ways of writing multiple states mean the same thing. Thus, numbers represented as $`n`$-qubit vectors lie in a space of dimension $`2^n`$, and may be written as $`1\times 2^n`$ column vectors (each of the $`2^n`$ slots in the column vector determined by the state of $`n`$-qubits), as illustrated for $`𝐇_2𝐇_2`$ above. We now introduce a matrix, $`W_{2^n}`$, that operates on these vectors. $`\mathrm{𝐓𝐡𝐞}\text{ }\mathrm{𝐖𝐚𝐥𝐬𝐡}\mathrm{𝐇𝐚𝐝𝐚𝐦𝐚𝐫𝐝}\text{ }\mathrm{𝐓𝐫𝐚𝐧𝐬𝐟𝐨𝐫𝐦𝐚𝐭𝐢𝐨𝐧}.\text{ }`$ The *Walsh-Hadamard transformation*, $`W_{2^n}`$, is defined recursively in the following way. Set $$W_2=H=\frac{1}{\sqrt{2}}\left(\begin{array}{cc}1& 1\\ 1& 1\end{array}\right),$$ (56) $$W_{2^n}=\frac{1}{\sqrt{2^n}}\left(\begin{array}{cc}W_{2^{n1}}& W_{2^{n1}}\\ W_{2^{n1}}& W_{2^{n1}}\end{array}\right),\text{ for }n>1.$$ (57) Note that $`W_4`$ is $$W_4=W_2W_2=\frac{1}{2}\left(\begin{array}{cc}1W_2& 1W_2\\ 1W_2& 1W_2\end{array}\right)=\frac{1}{2}\left(\begin{array}{cccc}1& 1& 1& 1\\ 1& 1& 1& 1\\ 1& 1& 1& 1\\ 1& 1& 1& 1\end{array}\right).$$ (58) Thus, for example $$W_4|uu=\frac{1}{2}\left(\begin{array}{cccc}1& 1& 1& 1\\ 1& 1& 1& 1\\ 1& 1& 1& 1\\ 1& 1& 1& 1\end{array}\right)\left(\begin{array}{c}1\\ 0\\ 0\\ 0\end{array}\right)=\frac{1}{2}\left(\begin{array}{c}1\\ 1\\ 1\\ 1\end{array}\right).$$ (59) We can rearrange the output, and see that it is a superposition of the elements of $`S=\{0,1,2,3\}`$: $`{\displaystyle \frac{1}{2}}\left(\begin{array}{c}1\\ 1\\ 1\\ 1\end{array}\right)={\displaystyle \frac{1}{2}}[\left(\begin{array}{c}1\\ 0\\ 0\\ 0\end{array}\right)+\left(\begin{array}{c}0\\ 1\\ 0\\ 0\end{array}\right)+\left(\begin{array}{c}0\\ 0\\ 1\\ 0\end{array}\right)+\left(\begin{array}{c}0\\ 0\\ 0\\ 1\end{array}\right)]={\displaystyle \frac{1}{2}}[|00+|01+|10+|11]`$ (80) $`={\displaystyle \frac{1}{2}}[|0+|1+|2+|3]={\displaystyle \frac{1}{\sqrt{2^n}}}{\displaystyle \underset{x=0}{\overset{2^n1}{}}}|x`$ (81) where here $`n=2`$, and we have mapped the binary numbers to their decimal equivalents. Thus, if $`|\psi =W_4|uu`$ and we take a measurement of $`|\psi `$, we will find a given number $`y`$, $`yS`$, with probability $`[\frac{1}{2}]^2=\frac{1}{4}`$. We may take the vectors $`|x`$ as basis vectors for our Hilbert space $`𝐇_4`$. Applying $`W_{2^n}`$ to $`n`$-bits, all in state $`|0`$, results in an equally weighted superposition of all states (numbers) in $`S=\{0,1,\mathrm{},2^n1\}`$: $$W_{2^n}|00\mathrm{}000=\frac{1}{\sqrt{2^n}}\underset{x=0}{\overset{2^n1}{}}|x.$$ (82) What happens if the qubits in the initial state of the quantum register are not all $`|0`$ (not all $`|u`$)? Define the *bit-wise inner product, or dot product, $`xy`$*, for $`x=x_{n1}x_{n2}\mathrm{}x_2x_1x_0`$, $`y=y_{n1}y_{n2}\mathrm{}y_2y_1y_0`$, as $`xy=x_{n1}y_{n1}+x_{n2}y_{n2}+\mathrm{}+x_2y_2+x_1y_1+x_0y_0`$ mod 2. (In the present example, taking the result mod 2 is redundant.) Then if the register was initially in state $`|y`$, the transformation is $$|\psi =W_{2^n}|y=\underset{x=0}{\overset{2^n1}{}}(1)^{xy}|x.$$ (83) For example, suppose $`|y`$ is the 3-qubit state $`|110`$. Then the bit-wise dot products and signs are shown in Table V. Thus we may write the output state $`|\psi `$ as $`|\psi =W_{2^n}|y={\displaystyle \frac{1}{\sqrt{2^3}}}(|000+|001|010|011|100|101+|110+|111)`$ (84) $`={\displaystyle \frac{1}{\sqrt{2^3}}}(|0+|1|2|3|4|5+|6+|7).`$ (85) The transformation of qubits must be *unitary*. Recall that a matrix $`U`$ is unitary if its inverse is equal to its complex conjugate transpose: $`U^1`$ = $`U^{}`$. Thus $`U^{}U=\mathrm{𝟏}`$. (For a Hermitian matrix $`M`$, $`M^{}=M`$, so a Hermitian matrix is unitary provided $`M^2=\mathrm{𝟏}`$.) The Pauli spin matrices, the Hadamard matrix $`H`$, and the Walsh matrix $`W_{2^n}`$ are all unitary. A unitary transformation conserves lengths of vectors. This can be seen if we compare the squared length of $`|\psi `$ and $`U|\psi `$: $`\psi |\psi =|\psi |^2`$ (86) $`\psi |U^{}U|\psi =\psi |\mathrm{𝟏}|\psi =|\psi |^2.`$ (87) One more unitary transformation we will need is the following: $$U_f|x|y=|x|y+_2f(x),$$ (88) where $`f:\{0,1\}\{0,1\}`$, and $`+_2`$ means addition modulo $`2`$. Note that $`U_f`$ operates on two qubits at once, $`|x|y`$. In this case, the $`|x`$ qubit is considered the *control* qubit and does not change in the operation; $`|y`$ is the data or *target* qubit, and changes according to whether $`f(x)=0`$ or $`f(x)=1`$. If $`f(x)=x`$, then $`U_f`$ here is called the *c-NOT* or *XOR* gate, often denoted by the negation symbol $`\neg `$. It takes the control and target qubits as inputs, and replaces the target qubit with the sum of the two inputs modulo 2: $$\neg |x|y=|x|y+_2x.$$ (89) Note for future reference with respect to the Grover search algorithm the effect of $`U_f`$ when $`|y=|0|1`$: $$U_f|x(|0|1)=|x[(|0|1)+_2f(x)].$$ (90) For $`f(x)=0`$ we have $$|x[(|0|1)+_2f(x)]=|x[|0|1]=|x(1)^{f(x)}(|0|1).$$ (91) For $`f(x)=1`$ we have $$|x[(|0|1)+_2f(x)]=|x[|1|0]=|x(1)^{f(x)}(|0|1).$$ (92) So, in summary, $$U_f|x(|0|1)=|x(1)^{f(x)}(|0|1).$$ (93) Note that if we modify the definition of $`f(x)`$ so that it is defined on the whole domain of $`S=\{0,1,2,\mathrm{},2^n1\}`$, $`f(x):xS\{0,1\}`$, then we can use $`f(x)`$ as an *indicator* or *characteristic* function, by letting $`f(a)=1`$ for some $`aS`$ and $`f(x)=0`$ for all $`xa`$. Denote this version of $`f(x)`$ as $`f_a(x)`$, and the associated unitary transformation as $`U_{f_a}|x|y=|x|y+_2f_a(x)`$. Then, as before, we have $$U_{f_a}|x(|0|1)=|x(1)^{f_a(x)}(|0|1).$$ (94) $`\mathrm{𝐓𝐡𝐞}\text{ }\mathrm{𝐆𝐫𝐨𝐯𝐞𝐫}\text{ }\mathrm{𝐒𝐞𝐚𝐫𝐜𝐡}\text{ }\mathrm{𝐀𝐥𝐠𝐨𝐫𝐢𝐭𝐡𝐦}.\text{ }`$ In computer science an *oracle* is a black box subroutine into which we are not allowed to look. An example of an oracle is our characteristic function $`f_a(x):xS\{0,1\}`$. It sets $`f_a(a)=1`$ and otherwise $`f_a(x)=0,\text{ }xa`$. If $`f_a(x)`$ is able to operate without our knowledge of what $`a`$ is, then $`f_a(x)`$ is an oracle. The values of $`x`$ may be an unsorted list—randomized telephone numbers for example (or ones which are sorted alphabetically by the owner’s names). The objective is to find $`a`$ by relying on the output of $`f_a(x)`$. If you had $`N=2^n`$ items, the expected number of queries to $`f_a(x)`$ to find $`a`$ with a probability of 50 percent would be $`\frac{N}{2}`$. Grover, however, showed a quantum computer could find the same item with a probability close to 100 percent in about $`\frac{\pi }{4}\sqrt{N}`$ searches. Suppose we are looking for the number $`a`$, where $`a`$ is $`n`$-bits. We will want to use our indicator function $`f_a(x)`$ as an oracle to help find $`a`$. *Initial Preparation.* First we prepare a qubit register with $`n+1`$ states, all of which are $`|0`$: $$|0|0\mathrm{}|0|0|0|0,$$ (95) where the tensor product has been explicitly written out for the right-most qubit to set it off from the rest. We apply the Walsh transform $`W_{2^n}`$ to the left $`n`$ $`|0`$ qubits and the simple transform $`H\sigma _x`$ to the last qubit. As we have seen before, $`|\psi _s=W_{2^n}|0|0\mathrm{}|0|0|0={\displaystyle \frac{1}{\sqrt{2^n}}}{\displaystyle \underset{x=0}{\overset{2^n1}{}}}|x`$ (96) $`H\sigma _x|0={\displaystyle \frac{1}{\sqrt{2}}}(|0|1),`$ (97) so that the state of the entire computer becomes $$|\psi _sH\sigma _x|0=\frac{1}{\sqrt{2^n}}\underset{x=0}{\overset{2^n1}{}}|x\frac{1}{\sqrt{2}}(|0|1).$$ (98) *Step One.* We then apply our unitary transformation $`U_{f_a}`$ $$U_{f_a}|x(|0|1)=|x(1)^{f_a(x)}(|0|1),$$ (99) to obtain $`U_{f_a}(|\psi _sH\sigma _x|0)={\displaystyle \frac{1}{\sqrt{2^n}}}{\displaystyle \underset{x=0}{\overset{2^n1}{}}}|x{\displaystyle \frac{1}{\sqrt{2}}}(1)^{f_a(x)}(|0|1)`$ (100) $`={\displaystyle \frac{1}{\sqrt{2^n}}}(1)^{f_a(x)}{\displaystyle \underset{x=0}{\overset{2^n1}{}}}|x{\displaystyle \frac{1}{\sqrt{2}}}(|0|1).`$ (101) The effect of $`U_{f_a}`$ is to change the sign on $`|x=|a`$ and to leave all the other superimposed states unchanged. You may ask, how did the sign $`(1)^{f_a(x)}`$ get transferred from the right-most qubit in equation (80) to the superposition of qubits in equation (81)? The answer is that the right-most qubit is allowed to *decohere*, to interact with the environment and to ‘collapse’ into $`|0`$ or $`|1`$. This forces the parameters that describe the bipartite state into the left $`n`$-qubit register. *Step Two.* Apply $`W_{2^n}`$ again to the left-most $`n`$ qubits. (Or apply $`W_{2^n}\mathrm{𝟏}_2`$ to $`n+1`$ qubits, where $`\mathrm{𝟏}_2`$ is the $`2\times 2`$ identity matrix.) *Step Three.* Let $`f_0(x)`$ be the indicator function for the state $`|x=|0`$. Apply $`U_{f_0}`$ to the current state of the qubit register (note the negation). This operation changes the sign on all states $`|x`$ except for $`|x=|0`$. That is, $`U_{f_0}`$ maps $`|0|0`$, and the negation of $`U_{f_0}`$, $`U_{f_0}`$ restores the original sign on $`|0`$ , but changes the sign on all other states. *Step Four.* Apply $`W_{2^n}`$ again to the left-most $`n`$ qubits. Repeat Steps One to Four $`\frac{\pi }{4}\sqrt{N}`$ times. Then sample the final state (the left-most $`n`$ qubits) $`|\psi _f`$. With close to probability 1, $`|\psi _f=|a`$. That’s the Grover search algorithm, but what does it mean? What do Steps One, Two, Three, and Four do? Short answer: they rotate the initial superposition $`|\psi _s`$ about the origin until it’s as close as possible to $`|a`$. Let’s see the details. Another way to think of $`U_{f_a}`$, in Step One, is as the matrix $`\mathrm{𝟏}2|aa|`$ operating on the left-most $`n`$ qubits. Applying this operation to $`|x`$ yields $`|x`$ for all basis states $`|x|a`$ but $`|x`$ for $`|x=|a`$. Similarly, another way to think of $`U_{f_0}`$, in Step Three, is as the matrix $`\mathrm{𝟏}2|00|`$. Applying this operation to $`|x`$ yields $`|x`$ for all basis states $`|x|0`$ but $`|0`$ for $`|x=|0`$. Step One is, geometrically, a reflection $`R_a`$ of $`|\psi _s`$ about the hyperplane orthogonal to $`|a`$ to a vector $`|\psi _s^R`$. Since $`W_{2^n}^2=\mathrm{𝟏}`$, Steps Two to Four correspond to $`W_{2^n}U_{f_0}W_{2^n}^1`$. The operation $`W_{2^n}U_{f_0}W_{2^n}^1`$ would correspond to a further reflection of $`|\psi _s^R`$ about the hyperplane orthogonal to the original $`|\psi _s=\frac{1}{\sqrt{2^n}}_{x=0}^{2^n1}|x`$. However, this isn’t what we want. Instead, let $`|\psi _s^{}`$ be a unit vector perpendicular to $`|\psi _s`$. The operation $`W_{2^n}U_{f_0}W_{2^n}^1`$ corresponds to a further reflection $`R_s`$ of $`|\psi _s^R`$ about the hyperplane orthogonal to $`|\psi _s^{}`$. Call this furtherly reflected vector $`|\psi _s^{^{}}`$. The net effect is a rotation $`R_sR_a=W_{2^n}U_{f_0}W_{2^n}^1U_{f_a}`$ of $`|\psi _s|\psi _s^{^{}}`$ in the plane spanned by $`|\psi _s`$ and $`|a`$. (By the plane spanned by $`|\psi _s`$ and $`|a`$ we mean all states of the form $`c|\psi _s+d|a`$, where $`c,d𝐂`$.) To summarize: Let $`\theta `$ be the angle between $`|\psi _s`$ and the unit vector orthogonal to $`|a`$, the latter designated $`|a^{}`$. For simplicity we assume a counter-clockwise ordering $`|a^{}`$, $`|\psi _s`$, $`|a`$. Then the combination $`R_sR_a`$ is a counter-clockwise rotation of $`|\psi _s`$ by $`2\theta `$, so that the angle between $`|a^{}`$ and $`|\psi _s`$ is now $`3\theta `$. That is, $`R_sR_a`$ moves $`|\psi _s`$ *away* from $`|a^{}`$, the vector orthogonal to $`|a`$, and hence moves $`|\psi _s`$ *toward* $`|a`$ itself by the angle $`2\theta `$. The whole idea of the Grover search algorithm is to rotate the state $`|\psi _s`$ about the origin, in the plane spanned by $`|\psi _s`$ and $`|a`$, until $`|\psi _s`$ is as close as possible to $`|a`$. Then a measurement of $`|\psi _s`$ will yield $`|a`$ with high probability. How much do we rotate (how many times do we apply $`R_sR_a`$)? We don’t want to overshoot or undershoot by rotating too much or too little. We want to rotate $`|\psi _s`$ around to $`|a`$ and then stop. Consider the vector or state $`|\psi _s`$ lying initially in the plane formed by $`|a^{}`$ and $`|a`$, with the angle between $`|\psi _s`$ and $`|a^{}`$ equal to $`\theta `$. That means we can write $`|\psi _s`$ as the initial superposition $$|\psi _s=cos\theta |a^{}+sin\theta |a.$$ (102) After $`k`$ applications of $`R_sR_a=W_{2^n}U_{f_0}W_{2^n}^1U_{f_a}`$, the state is $$(R_sR_a)^k|\psi _s=cos(2k+1)\theta |a^{}+sin(2k+1)\theta |a.$$ (103) Note that if $`(2k+1)\theta =\frac{\pi }{2}`$, then $`cos(2k+1)\theta =0`$, $`sin(2k+1)\theta =1`$, so that $$(R_sR_a)^k|\psi _s=|a.$$ (104) Now this may not be achievable, because $`k`$ must be a whole number, but let’s solve for the closest integer, where $`[]_{nint}`$ denotes nearest integer: $$k=[\frac{\pi }{4\theta }\frac{1}{2}]_{nint}.$$ (105) Remember that the inner product of two unit vectors gives the cosine of the angle between them, and that the *initial* angle between $`|a`$ and $`|\psi _s`$ is $`\frac{\pi }{2}\theta `$. Therefore $$a|\psi _s=\frac{1}{\sqrt{2^n}}=cos(\frac{\pi }{2}\theta )=sin(\theta ).$$ (106) For $`N=2^n`$ large, we can set $`sin\text{ }\theta \theta `$. Thus, substituting $`\frac{1}{\sqrt{N}}=\theta `$ into our equation for $`k`$, we obtain $$k=[\frac{\pi }{4}\sqrt{N}\frac{1}{2}]_{nint}.$$ (107) This value of $`k`$, then, obtains $`(R_sR_a)^k|\psi _s=|a`$ with probability close to $`1`$. $`\mathrm{𝐆𝐫𝐨𝐯𝐞𝐫}\text{ }\mathrm{𝐬𝐞𝐚𝐫𝐜𝐡}\text{ }\mathrm{𝐞𝐱𝐚𝐦𝐩𝐥𝐞}.\text{ }`$ Here is an example of Grover search for $`n=3`$ qubits, where $`N=2^n=8`$. (We omit reference to qubit $`n+1`$, which is in state $`\frac{1}{\sqrt{2}}(|0|1)`$ and does not change. The dimension of the unitary operators for this example is thus $`2^n=8`$ also.) Suppose the unknown number is $`|a=|5`$. The matrix or black box oracle $`U_{f_a}`$ is then $$U_{f_5}=\left(\begin{array}{cccccccc}1& 0& 0& 0& 0& 0& 0& 0\\ 0& 1& 0& 0& 0& 0& 0& 0\\ 0& 0& 1& 0& 0& 0& 0& 0\\ 0& 0& 0& 1& 0& 0& 0& 0\\ 0& 0& 0& 0& 1& 0& 0& 0\\ 0& 0& 0& 0& 0& 1& 0& 0\\ 0& 0& 0& 0& 0& 0& 1& 0\\ 0& 0& 0& 0& 0& 0& 0& 1\end{array}\right).$$ (108) (Remember that numbering starts with 0 and ends with 7, so that the -1 here is in the slot for $`|5`$.) This matrix reverses the sign on state $`|5`$, and leaves the other states unchanged. The Walsh matrix $`W_8`$ is $$W_8=\frac{1}{\sqrt{2^3}}\left(\begin{array}{cccccccc}1& 1& 1& 1& 1& 1& 1& 1\\ 1& 1& 1& 1& 1& 1& 1& 1\\ 1& 1& 1& 1& 1& 1& 1& 1\\ 1& 1& 1& 1& 1& 1& 1& 1\\ 1& 1& 1& 1& 1& 1& 1& 1\\ 1& 1& 1& 1& 1& 1& 1& 1\\ 1& 1& 1& 1& 1& 1& 1& 1\\ 1& 1& 1& 1& 1& 1& 1& 1\end{array}\right).$$ (109) The matrix $`U_{f_0}`$ is $$U_{f_0}=\left(\begin{array}{cccccccc}1& 0& 0& 0& 0& 0& 0& 0\\ 0& 1& 0& 0& 0& 0& 0& 0\\ 0& 0& 1& 0& 0& 0& 0& 0\\ 0& 0& 0& 1& 0& 0& 0& 0\\ 0& 0& 0& 0& 1& 0& 0& 0\\ 0& 0& 0& 0& 0& 1& 0& 0\\ 0& 0& 0& 0& 0& 0& 1& 0\\ 0& 0& 0& 0& 0& 0& 0& 1\end{array}\right).$$ (110) This matrix changes the sign on all states except $`|0`$. Finally, we have the repeated step $`R_sR_a`$ in the Grover algorithm: $$R_sR_5=W_8U_{f_0}W_8^1U_{f_5}=\frac{1}{4}\left(\begin{array}{cccccccc}3& 1& 1& 1& 1& 1& 1& 1\\ 1& 3& 1& 1& 1& 1& 1& 1\\ 1& 1& 3& 1& 1& 1& 1& 1\\ 1& 1& 1& 3& 1& 1& 1& 1\\ 1& 1& 1& 1& 3& 1& 1& 1\\ 1& 1& 1& 1& 1& 3& 1& 1\\ 1& 1& 1& 1& 1& 1& 3& 1\\ 1& 1& 1& 1& 1& 1& 1& 3\end{array}\right).$$ (111) The *initial preparation* is $$W_8|0|0|0=\frac{1}{\sqrt{2^3}}\left(\begin{array}{c}1\\ 1\\ 1\\ 1\\ 1\\ 1\\ 1\\ 1\end{array}\right).$$ (112) Since $`N=2^3=8`$ we calculate the number of rotations $`k`$ as the nearest integer: $$k=[\frac{\pi }{4}\sqrt{8}\frac{1}{2}]_{nint}=2.$$ (113) Thus, after the first rotation, the state becomes $$R_sR_5W_8|0|0|0=\frac{1}{4\sqrt{2}}\left(\begin{array}{c}1\\ 1\\ 1\\ 1\\ 1\\ 5\\ 1\\ 1\end{array}\right)$$ (114) and, after the second rotation, $$(R_sR_5)^2W_8|0|0|0=\frac{1}{8\sqrt{2}}\left(\begin{array}{c}1\\ 1\\ 1\\ 1\\ 1\\ 11\\ 1\\ 1\end{array}\right).$$ (115) Note that the amplitude for $`|5`$ is now $`\frac{11}{8\sqrt{2}}`$. A measurement of $`(R_sR_5)^2W_8|0|0|0`$ will thus yield $`|5`$ with probability $`(\frac{11}{8\sqrt{2}})^2=.9453`$. $`\mathrm{𝐓𝐡𝐞}\text{ }\mathrm{𝐠𝐮𝐞𝐬𝐬}\text{ }𝐚\text{ }\mathrm{𝐧𝐮𝐦𝐛𝐞𝐫}\text{ }\mathrm{𝐠𝐚𝐦𝐞}\text{ }𝐈.\text{ }`$ Bob challenges Alice to the following game. Alice is to chose a number $`a`$ from $`S=\{0,1,\mathrm{},N1\}`$, and he is to attempt to guess it, with a certain number of tries $`k`$. Alice acts as the oracle $`U_{f_a}`$ after each of Bob’s turns. They agree on $`N=2^{30}=1,073,741,824`$. Alice knows that, classically, Bob will require $`\frac{N}{2}=2^{29}=536,870,912`$ tries to guess the number with a probability of 50 percent, so she agrees with Bob to allow up to $`k=100,000,000`$, believing that the advantage is all hers. Bob, however, intends to use the Grover search algorithm, and never intends to guess more than $`k=[\frac{\pi }{4}\sqrt{2^{30}}\frac{1}{2}]_{nint}=25,735`$ times. Bob initially sets up $`N+1`$ qubits as $$|\psi _sH\sigma _x|0=\frac{1}{\sqrt{2^n}}\underset{x=0}{\overset{2^n1}{}}|x\frac{1}{\sqrt{2}}(|0|1),$$ (116) as in equation (78). He presents the left-most $`n`$ qubits, $`|\psi _s`$, to Alice. This is followed by Alice’s move of $`R_a`$, followed by Bob’s play of $`R_s`$, and so on, until after $`k`$ moves the state of the $`n`$-qubit system is: $$(R_sR_a)^k|\psi _s=cos(2k+1)\theta |a^{}+sin(2k+1)\theta |a.$$ (117) The system is then measured and Bob wins with a probability of $`|sin(2k+1)\theta |^2`$. To Alice’s surprise she finds that Bob wins repeatedly, despite playing only a small number of his allowed moves. (Bob’s probability of winning is $`p1\frac{1}{N}`$.) After a number of games she realizes Bob always plays the same number of moves $`k=25,735`$. She becomes suspicious that there is some conspiracy afoot. $`\mathrm{𝐓𝐡𝐞}\text{ }\mathrm{𝐁𝐞𝐫𝐧𝐬𝐭𝐞𝐢𝐧}\mathrm{𝐕𝐚𝐳𝐢𝐫𝐚𝐧𝐢}\text{ }\mathrm{𝐨𝐫𝐚𝐜𝐥𝐞}.\text{ }`$ Previously we defined the bitwise inner product $`xy`$. Let’s substitute for $`y`$ a constant vector $`a`$ of 0s and 1s, and let $`f_{bv}^a:\{0,1\}^n\{0,1\}`$ be defined as $$f_{bv}^a(x,a)=xa$$ (118) with an associated transform $$T_{bv}^a|x=(1)^{f_{bv}^a}|x=(1)^{xa}|x.$$ (119) This is the Bernstein-Vazirani oracle. How many measurements of $`f_{bv}^a(x,a)`$ would be required to find $`a`$? Classically you would have to perform measurements for all possible values of $`x`$, and then solve a set of linear equations for $`a`$. But quantum mechanically solving for $`a`$ only takes one step. To see why, refer back to equation (63) and the calculation in Table V for the Walsh transform of an initial state $`|y|0`$. Now compare the effect of the transform $`T_{bv}^a`$ on an equal superposition of all states: $$T_{bv}^a|\psi _s=\frac{1}{\sqrt{2^n}}\underset{x=0}{\overset{2^n1}{}}T_{bv}^a|x=\frac{1}{\sqrt{2^n}}\underset{x=0}{\overset{2^n1}{}}(1)^{xa}|x.$$ (120) This is just the Walsh transform of an initial state $`|a`$! Therefore we can find $`|a`$ with another application of the Walsh transform (which is its own inverse): $$W_{2^n}T_{bv}^a|\psi _s=|a.$$ (121) $`\mathrm{𝐓𝐡𝐞}\text{ }\mathrm{𝐠𝐮𝐞𝐬𝐬}\text{ }𝐚\text{ }\mathrm{𝐧𝐮𝐦𝐛𝐞𝐫}\text{ }\mathrm{𝐠𝐚𝐦𝐞}\text{ }\mathrm{𝐈𝐈}.\text{ }`$ Alice says to Bob, you are getting too many guesses. Either change the game or I won’t play anymore. Bob says: I don’t know why you are complaining. I’m only making a tiny fraction of the number of guesses we agreed on. But I’ll tell you what. I will make only *two* guesses–a preliminary guess, you will give me some feedback information, and then I will make a second and final guess of the number. The feedback I need is $`T_{bv}^a`$ applied as an oracle to my initial guess. (Of course Bob plans to submit $`|\psi _s`$ as his initial guess.) Alice agrees, and the game proceeds as follows: Bob: $`prepares\text{ }|\psi _s=W_{2^n}|0\mathrm{}00=\frac{1}{\sqrt{2^n}}_{x=0}^{2^n1}|x`$ Alice: $`T_{bv}^a|\psi _s=\frac{1}{\sqrt{2^n}}_{x=0}^{2^n1}(1)^{xa}|x`$ Bob: $`W_{2^n}T_{bv}^a|\psi _s=|a`$ . Bob wins. Again, the key feature was the ability to present a superposition of states to Alice’s oracle. ## .7 Shor’s factoring algorithm Shor’s algorithm is a key result in quantum computation, so we want to look at it in some modest detail. It will form the basis of the RSA game. We will need as preliminaries Euler’s theorem and the quantum Fourier transform $`F`$. $`\mathrm{𝐄𝐮𝐥𝐞𝐫}^{}𝐬\text{ }\mathrm{𝐭𝐡𝐞𝐨𝐫𝐞𝐦}.`$ Let $`N`$ be an integer, and let $`a`$ be an integer less than $`N`$ and relatively prime to $`N`$. Euler’s theorem (OO, , chap. 12) says that $$a^\varphi =1\text{ }mod\text{ }N.$$ (122) Here $`\varphi `$ is Euler’s totient function, and is the total number of integers less than $`N`$ that are relatively prime to $`N`$. Example: Let N = 77. In this case $`\varphi =60`$, so $`23^{60}=1`$ mod $`77`$, $`39^{60}=1`$ mod $`77`$, etc. Euler’s theorem implies that the powers of any number relatively prime to $`N`$ cycle mod $`N`$: $$a,a^2,a^3,\mathrm{},a^{\varphi 1},a^\varphi =1,a,a^2,a^3,\mathrm{}.$$ (123) Thus $`\varphi `$ is the maximum length of a cycle or period. Of course, for a given $`a`$, there may be a smaller $`s<\varphi `$ such that $`a^s=1`$ mod $`N`$. But in that case it is clear $`s`$ divides $`\varphi `$. The smallest value of $`s`$ such that $`a^s`$ = 1 mod N is called the *order* of $`a`$, which in the Shor algorithm below we denote by $`r`$. Given knowledge of $`\varphi `$, or any $`s`$ or $`r`$ for a given $`a`$, we can factor $`N`$. Since $`a^\varphi =1`$ mod $`N`$, we have, for even $`\varphi `$, $`(a^{\frac{\varphi }{2}}+1)(a^{\frac{\varphi }{2}}1)=0`$ mod $`N`$. Let $`gcd(x,y)`$ denote the greatest common divisor of $`x`$ and $`y`$. We then check $`gcd(N,a^{\frac{\varphi }{2}}+1)`$ and $`gcd(N,a^{\frac{\varphi }{2}}1)`$ for a factor. If we don’t get a factor, we divide $`\varphi `$ again by two (if the previous division left an even exponent), or else try another value for $`a`$. Example: Let $`N=77`$, and $`a=2`$. We find that $`2^{60}=1`$ mod $`77`$, and upon division of $`\varphi `$ by $`2`$, also $`2^{30}=1`$ mod $`77`$. Hence we look at $`2^{15}`$ mod $`N=43`$. We find that $`gcd(77,44)=11`$ and $`gcd(77,42)=7`$. These are the two factors of 77. Obviously, this is not the best way to factor a number, normally, but it is ideally suited for a quantum algorithm. $`\mathrm{𝐐𝐮𝐚𝐧𝐭𝐮𝐦}\text{ }\mathrm{𝐅𝐨𝐮𝐫𝐢𝐞𝐫}\text{ }\mathrm{𝐭𝐫𝐚𝐧𝐬𝐟𝐨𝐫𝐦}.`$ The quantum Fourier transform looks a lot like the discrete Fourier transform. For a given state $`|y`$ the quantum Fourier transform is the unitary transformation $$F|y=\frac{1}{\sqrt{2^n}}\underset{x=0}{\overset{2^n1}{}}e^{2\pi ixy/2^n}|x.$$ (124) In this definition, the term $`xy`$ denotes ordinary multiplication. It is *not* the bitwise dot product $`xy`$. Rather, if $`|x=7`$ and $`|y=6`$, then $`xy=42`$. (By contrast, the dot product is $`xy=76`$ mod $`2=111110`$ mod $`2=2`$ mod $`2=0`$.) $`F|y`$ is periodic in $`xy`$ with period $`2^n`$. The Hadamard matrix $`H`$ we saw previously is simply the Fourier transform for $`n=1`$. To see this, let x, y each be 0 or 1 in the term $$\frac{1}{\sqrt{2^n}}e^{2\pi ixy/2^n}$$ (125) where $`n=1`$. We obtain the matrix $$\frac{1}{\sqrt{2}}\left(\begin{array}{cc}e^0& e^0\\ e^0& e^{\pi i}\end{array}\right)=\frac{1}{\sqrt{2}}\left(\begin{array}{cc}1& 1\\ 1& 1\end{array}\right),$$ (126) remembering that $`e^{\pi i}=cos(\pi )+i\text{ }sin(\pi )=1+0=1`$. The inverse quantum Fourier transform $`F^1`$ simply reverses the sign on $`i`$: $$F^1|y=\frac{1}{\sqrt{2^n}}\underset{x=0}{\overset{2^n1}{}}e^{2\pi ixy/2^n}|x.$$ (127) $`\mathrm{𝐒𝐡𝐨𝐫}^{}𝐬\text{ }\mathrm{𝐟𝐚𝐜𝐭𝐨𝐫𝐢𝐧𝐠}\text{ }\mathrm{𝐚𝐥𝐠𝐨𝐫𝐢𝐭𝐡𝐦}.\text{ }`$We want to find a factor of a number $`N`$, where $`2^{2n2}<N^2<2^{2n}`$. Shor’s factoring algorithm on a quantum computer runs in $`O((log\text{ }N)^3)`$ steps. We need a quantum computer with two registers (which we shall refer to simply as left and right). The left register contains $`2n`$ qubits, and the right register contains $`log__2N`$ qubits. The values of the qubits in both registers are initialized to $`|0`$: $$|00\mathrm{}0|00\mathrm{}0.$$ (128) *Step 1:* Chose $`m`$, $`2mN2`$. If $`gcd(m,N)2`$, we have found a proper factor of $`N`$. Otherwise proceed as follows, in Steps 2-5. *Step 2:* Do a Walsh transform $`W_{2^{2n}}`$ of the qubits in the left register to create a superposition of all states in the left register: $$(W_{2^{2n}}\mathrm{𝟏}_{log_2N})(|00\mathrm{}0|00\mathrm{}0)=|\psi _s|00\mathrm{}0=\frac{1}{\sqrt{2^{2n}}}\underset{x=0}{\overset{2^{2n}1}{}}|x|00\mathrm{}0.$$ (129) *Step 3:* Apply the transform $`f_m(|x|00\mathrm{}0)|x|m^x\text{ mod }N`$: $$f_m(|\psi _s|00\mathrm{}0)=\frac{1}{\sqrt{2^{2n}}}\underset{x=0}{\overset{2^{2n}1}{}}|x|m^x\text{ mod }N.$$ (130) Note that at this point, if we measured the right register, or allowed it to decohere, it would collapse into a given value of $`m^x`$ mod $`N`$, such as $`Z=m^z`$ mod $`N`$. Hence, in the left register, all amplitudes of states would go to zero, except for those states $`x`$ such that $`m^x`$ mod $`N=Z`$. If, for example, the order of $`m`$ was $`5`$, then the amplitudes of states would read something like: $$\mathrm{},0,0,0,c,0,0,0,0,c,0,0,0,0,c,0,0,0,0,c,0,0\mathrm{}$$ (131) The amplitude would be non-zero on every $`5th`$ value. The states were previously in an equal superposition with amplitude $`\frac{1}{\sqrt{2^{2n}}}`$, but the surviving values would now have amplitude approximately $`c=\frac{1}{\sqrt{\frac{2^{2n}}{5}}}`$. This is the idea, although (following Shor), we don’t actually observe the right register at this point. Instead we proceed to Step 4. *Step 4:* Do a quantum Fourier transform $`F`$ on the qubits in the left register: $$(F\mathrm{𝟏})(f_m(|\psi _s|00\mathrm{}0)=\frac{1}{2^{2n}}\underset{x=0}{\overset{2^{2n}1}{}}\underset{y=0}{\overset{2^{2n}1}{}}e^{2\pi ixy/2^{2n}}|y|m^x\text{ mod }N.$$ (132) *Step 5:* Observe the system registers. This will give some concrete value of $`w`$ for $`y`$ and $`m^z`$ mod $`N`$ for $`m^x`$ mod $`N`$: $$(F\mathrm{𝟏})(f_m(|\psi _s|00\mathrm{}0)|w,m^z\text{ mod }N$$ (133) with probability equal to the square of the associated amplitude: $$|\frac{1}{2^{2n}}\underset{x:m^x=m^z\text{ mod }N}{}e^{2\pi ixw/2^{2n}}|^2.$$ (134) Thus with high probability, the observed $`w`$ will be near an integer multiple of $`\frac{2^{2n}}{r}`$. This ends the quantum part of the calculation. We now use the result to determine the period $`r`$. First find the fraction that best approximates $`\frac{w}{2^{2n}}`$ with denominator $`r^{}<N<2^n`$: $$|\frac{w}{2^{2n}}\frac{d^{}}{r^{}}|<\frac{1}{2^{2n+1}}.$$ (135) This may be done using continued fractions (see (HW, , chapter 12)). Second try $`r^{}`$ in the role of $`r`$. If $`m^r^{}=1`$ mod $`N`$, we have, for even $`r^{}`$, $`(m^{\frac{r^{}}{2}}1)(m^{\frac{r^{}}{2}}+1)=0\text{ mod }N`$. We then check $`gcd(N,m^{\frac{r^{}}{2}}1)`$ and $`gcd(N,m^{\frac{r^{}}{2}}+1)`$ for a factor of N. In the event $`r^{}`$ is odd, or if $`r^{}`$ is even and we don’t obtain a factor, we repeat the steps $`O(`$log log $`N)`$ times using the same value for $`m`$. If that doesn’t work, we change $`m`$ and start over. ## .8 The RSA game RSA is an encryption system widely used in banking and elsewhere. Consider the ring of integers $`Z_N`$, where $`N=pq`$ for two distinct large primes $`p`$ and $`q`$. For encryption, RSA allows only the *units* of $`Z_N`$ (i.e., eliminate all multiples of $`p`$ or $`q`$ from $`Z_N`$). The remaining set of integers, called $`Z_N^{}`$, is an abelian group under multiplication, with order (Euler’s totient function) $`\varphi =(p1)(q1)=(n+1)(p+q)`$. The RSA crypto system choses a relatively small odd integer $`e`$, and calculates $`d=e^1`$ mod $`\varphi `$. A message $`M`$ in $`Z_n^{}`$ is then encrypted as $`M^e`$ mod $`N`$, and decrypted as $`M^{ed}=M^{\varphi +1}=M`$ mod $`N`$. The numbers $`e`$ and $`N`$ are publicly known, while the decryption key $`d`$ is known only to the message recipient. Alice challenges Bob to the following game. She will create a public key $`N`$ and $`e`$, and encrypt a message $`M`$. The three components $`(N,e,M^e)`$ will be sent to Bob. If Bob can decrypt the message, $`M^eM`$, within (log $`N)^3`$ steps, Bob wins $1,000. Else he loses $1,000. Now RSA uses very large numbers $`N`$. But we are going to use an extremely simple example in order to illustrate the steps in Shor’s algorithm. We assume that Alice sends Bob the triplet $`(77,11,67)`$. We first note that $`77^2=5929`$, and $`2^{12}<5929<2^{14}`$. The left quantum register will need 14 qubits, while the right register will require 7 qubits. *Step 1:* Bob randomly chooses $`m=39`$, where $`23975`$. The $`gcd(39,77)=1`$, so Bob proceeds to Step 2. *Step 2:* In the left qubit register, Bob creates a superposition of all numbers from 0 to $`16383=2^{14}1`$. *Step 3:* Bob applies the transform $`f_m`$ which associates to each $`x`$ in the superposition, the value $`39^x`$ mod $`77`$. Since $`39^{30}`$ mod $`77=1`$, we have $`m^x=1`$ mod $`77`$, for $`xS=\{30,60,90,120,150,\mathrm{},16380\}`$. That is $`m=39`$ has period $`r=30`$. But Bob doesn’t know this yet. *Step 4:* Bob does a quantum Fourier transform on the left register, which contains the values of $`x`$. He then observes both registers and gets $`w=14,770`$ for the left register state, and $`Z=53`$ for the value of $`39^z`$ mod $`77`$ in the right register. Bob now wants to find the fraction that best approximates $`\frac{14770}{16384}`$ with denominator less than $`77`$. This fraction is very close to $`\frac{27}{30}`$, so Bob tries $`r^{}=30`$, or $`\frac{r^{}}{2}=15`$. He gets $`39^{15}1`$ mod $`77=42`$, $`39^{15}+1`$ mod $`77=44`$, and $`gcd(77,42)=7`$, $`gcd(77,44)=11`$. With these two factors in hand, Bob calculates $`\varphi =(71)(111)=60`$. Therefore for the decryption key $`d`$, he wants $`d=e^1`$ mod $`60`$, which gives $`d=11^1`$ mod $`60=11`$. The decryption key is the same as the encryption key. (This is only a result of the trivially small modulus $`N=77`$ we used.) Bob now decrypts Alice’s encrypted message $`(M^e)^d=67^{11}`$ mod $`77=23`$. Bob tells Alice the message $`M=23`$ and collects his $1,000. ## .9 Nash equilibrium and prisoner’s dilemma We want to look at $`2\times 2`$ games that are not zero sum, and the traditional game theoretic concept of Nash equilibrium, and to extend it to quantum games. Both Alice and Bob may gain from a game, but may or may not do as well as some obtainable maximum. We assume both try to maximize utility, or *expected utility* with mixed strategies or uncertain outcomes, and that utility can be assigned a cardinal number PF . Non-zero sum games are traditionally presented in static form. A matrix of payoffs corresponding to moves is given, and some notion of *equilibrium* is presented, without explaining how the players got to that point. But once they get there, they are expected to stay. That’s because they have a *dominant strategy* that indicates they are better off playing the corresponding move. Let $`s_A^iS_A`$ be moves (including convex combinations of simple moves, if appropriate) available to Alice, and $`s_B^jS_B`$ be moves available to Bob. Then a *dominant strategy* for Alice is a move $`s_A`$ such that the payoff $`\pi _A`$ to Alice has the property $$\pi _A(s_A,s_B^j)\pi _A(s_A^i,s_B^j)$$ (136) for all $`s_A^iS_A`$, $`s_B^jS_B`$, provided such a move exists. For an example, consider Table VI. Alice and Bob each have two possible moves, labeled C (cooperate) or D (defect). The values in parenthesis represent the payoffs $`\pi `$; the first number is the payoff to Alice, the second number is the payoff to Bob. Clearly for Alice $`s_A=D`$, because if Bob plays $`C`$, $`\pi _A(D,C)=5>3`$, while if Bob plays $`D`$, $`\pi _A(D,D)=1>0`$. For similar reasons, $`s_B=D`$ also, so the game will be in *equilibrium* with $`\{s_A,s_B\}=\{D,D\}`$ and $`\{\pi (s_A),\pi (s_B)\}=\{1,1\}`$. This outcome is referred to as *Prisoner’s Dilemma* because clearly Bob and Alice would each be better off if both played C, which would yield $`\pi _A=\pi _B=3`$. A *Nash equilibrium* is a combination of moves $`\{s_A,s_B\}`$ such that neither party can increase his or her payoff by unilaterally departing from the given equilibrium point: $`\pi _A(s_A,s_B)\pi _A(s_A^i,s_B),`$ (137) $`\pi _B(s_A,s_B)\pi _B(s_A,s_B^j).`$ (138) In Table VI, $`\{D,D\}`$, yielding payoffs $`\{1,1\}`$ is a Nash equilibrium, because if Alice switches to C, her payoff goes from 1 to 0, and similarly for Bob. A payoff point $`\{\pi _A,\pi _B\}`$ is *jointly dominated* by a different point $`\{\pi _A^{},\pi _B^{}\}`$ if $`\pi _A^{}\pi _A`$ and $`\pi _B^{}\pi _B`$, and one of the inequalities is strict. In Table VI, the point $`\{1,1\}`$ is jointly dominated by $`\{3,3\}`$. A pair of payoffs $`\{\pi _A,\pi _B\}`$ is *Pareto optimal* if it is not jointly dominated by another point, and if neither party can increase his or her payoff without decreasing the payoff to the other party. In Table VI, the point $`\{3,3\}`$ is Pareto optimal, because unilateral departure from it by either Alice or Bob decreases the payoff to the other party. What about $`\{1,1\}`$? Here, too, neither party can increase their payoff without decreasing the payoff to the other party (indeed, neither can unilaterally increase his payoff at all). However, $`\{1,1\}`$ is jointly dominated by $`\{3,3\}`$, so it is not Pareto optimal. An *evolutionarily stable strategy* (ESS) is a more restrictive notion than Nash equilibrium. (That is, strategies that are evolutionarily stable form a subset of Nash equilibria.) Strategy $`s_i`$ is *evolutionarily stable* against $`s_j`$ if $`s_i`$ performs better than $`s_j`$ against $`s_i+(1\eta )s_j`$ for sufficiently small $`\eta `$. The notion is that of a population playing $`s_i`$ that is invaded by mutants playing $`s_j`$. An ESS is then defined as a strategy that is evolutionarily stable against all other strategies. Note that an ESS holds for $`\eta `$ sufficiently small, say $`\eta [0,\eta _0)`$. The value $`\eta _0`$ is called the *invasion barrier*. For values of $`\eta >\eta _0`$, $`s_i`$ no longer performs better than $`s_j`$ against the combination, so members of the population will switch to $`s_j`$. We will return to this concept in the *evolutionarily stable strategy game* considered later. ## .10 Escaping prisoner’s dilemma in a quantum game We now have enough background to tentatively define a quantum game. A *quantum game $`\mathrm{\Gamma }`$* is an interaction between two or more players with the following elements: $`\mathrm{\Gamma }=\mathrm{\Gamma }(𝐇,\mathrm{\Lambda },\{s_i\}_j,\{\pi _i\}_j)`$. $`𝐇`$ is a Hilbert space, $`\mathrm{\Lambda }`$ represents the initial state of the game, $`\{s_i\}_j`$ is the set of moves of player $`j`$, while $`\{\pi _i\}_j`$ is a set of payoffs to player $`j`$. The object of the game is that of endogenously determining the strategies that maximize the payoffs to player $`j`$. In the course of doing so, we may or may not determine an equilibrium to the game, and the value $`\overline{\pi }_j`$ of the game to player $`j`$. We want, at this point, to give an introduction to the *quantum* version of Prisoner’s Dilemma, even though final details will be deferred until later. In the quantum version of prisoner’s dilemma EWL , each of Alice and Bob possesses a qubit and is able to perform manipulations on his/her own qubit. Each qubit lies in $`𝐇_2`$ which has as basis vectors $`|C`$ and $`|D`$, and the game lies in $`𝐇_2𝐇_2`$ with basis vectors $`|CC`$, $`|CD`$, $`|DC`$, and $`|DD`$. Alice’s qubit is the left-most qubit in each pair, while Bob’s is the right-most. The game is a simple quantum network. The initial state $`\mathrm{\Lambda }`$ of the game is $$\mathrm{\Lambda }=U|CC,$$ (139) where $`U`$ is a unitary operator, known both to Alice and Bob, that operates on both qubits. Alice and Bob have as strategic moves $`s_A`$, $`s_B`$, $`s_A=U_A`$ (140) $`s_B=U_B`$ (141) where $`U_A`$ and $`U_B`$ are unitary matrices that operate only on the respective player’s qubit. After Alice and Bob have made their moves, the state of the game is $$(U_AU_B)U|CC.$$ (142) Alice and Bob forward their qubits for final measurement. The inverse of the unitary operator $`U`$ is now applied, to bring the game to the state: $$U^{}(U_AU_B)U|CC.$$ (143) The measurement is then taken, and yields one of the four basis vectors of $`𝐇_2𝐇_2`$. The associated payoff values to Alice and Bob are those previously given in Table VI. How Alice and Bob escape prisoner’s dilemma in this quantum game by selection of their respective unitary matrices $`U_A`$, $`U_B`$ depends on their playing *entanglement*-related strategies. Therefore we will defer further discussion of the quantum prisoner’s dilemma game until we have considered entanglement in the next section. However, we wanted to make the point that *a pure quantum strategy is a unitary operator acting on the player’s qubit*. ## .11 Entanglement We have been considering vectors $`|\psi `$ in a Hilbert space $`𝐇`$. The vector or state $`|\psi `$ is *entangled* if it does not factor relative to a given tensor product decomposition of the Hilbert space, $`𝐇=𝐇_1𝐇_2`$. For example, the state $`|\psi _1=a|00+b|01`$ can be decomposed into a tensor product $$|\psi _1=a|00+b|01=|0(a|0+b|1),$$ (144) so it is not entangled. On the other hand, the state $`|\psi _2=a|00+b|11`$ cannot be decomposed into a tensor product, and is therefore entangled. Entangled states act as a single whole without reference to space or time. Any operation performed on one entangled qubit instantly affects the states of the qubits with which it is entangled. Entanglement generates ‘spooky action at a distance’. Instead of the orthonormal computational basis we have been using for Hilbert space, sometimes a different orthonormal basis, called the *Bell basis*, is used. The Bell basis is a set of maximally entangled states. For two-qubits in $`𝐇_4`$, we can denote this entangled basis as $`|b_0={\displaystyle \frac{1}{\sqrt{2}}}(|00+|11)`$ (145) $`|b_1={\displaystyle \frac{1}{\sqrt{2}}}(|01+|10)`$ (146) $`|b_2={\displaystyle \frac{1}{\sqrt{2}}}(|00|11)`$ (147) $`|b_3={\displaystyle \frac{1}{\sqrt{2}}}(|01|10).`$ (148) It is easy to transform the computational basis into the Bell basis by using a combination of a Hadamard transformation $`H`$ and a c-NOT gate. First apply the Hadamard transform to the left-most qubit. Then apply c-NOT (review equation 69) with the left qubit as the source and the right qubit as the target. Shorthand for this transformation is $`\neg (H\mathrm{𝟏})`$: $`\neg (H\mathrm{𝟏})|00\neg {\displaystyle \frac{1}{\sqrt{2}}}(|0+|1)|0|b_0`$ (149) $`\neg (H\mathrm{𝟏})|01\neg {\displaystyle \frac{1}{\sqrt{2}}}(|0+|1)|1|b_1`$ (150) $`\neg (H\mathrm{𝟏})|10\neg {\displaystyle \frac{1}{\sqrt{2}}}(|0|1)|0|b_2`$ (151) $`\neg (H\mathrm{𝟏})|11\neg {\displaystyle \frac{1}{\sqrt{2}}}(|0|1)|1|b_3.`$ (152) We will now show how quantum entanglement can get players out of prisoner’s dilemma. ## .12 Return to the quantum Prisoner’s Dilemma Let’s return to the quantum version of Prisoner’s Dilemma. For consistency of notation, we map $`|C|0`$ and $`|D|1`$. When we left the final state of the game, equation (123), it had the form $$|\psi _f=U^{}(U_AU_B)U|00.$$ (153) When a measurement of the system is taken, it is projected into one of the four basis vectors $`|00`$, $`|01`$, $`|10`$, $`|11`$, with associated probability, yielding as expected payoff $`\overline{\pi }_A`$ to Alice (refer to Table VI): $$\overline{\pi }_A=3|\psi _f|00|^2+0|\psi _f|01|^2+5|\psi _f|10|^2+1|\psi _f|11|^2.$$ (154) The payoff probabilities depend on the final state of the game, which in turn depends on the unitary matrix $`U`$ and the player moves $`U_A`$ and $`U_B`$. Let’s consider each of these in turn. The purpose of the unitary matrix $`U`$ is to entangle Alice’s and Bob’s qubits. Without this entanglement the payoffs to Bob and Alice remain the same as in the classical game (namely, the Nash equilibrium of (1,1)). Let’s let our unitary matrix $`U`$ be (where $`n`$ simply means the tensor product $`n`$ times): $$U=\frac{1}{\sqrt{2}}(\mathrm{𝟏}^2+i\sigma _x^2).$$ (155) The inverse is $$U^{}=\frac{1}{\sqrt{2}}(\mathrm{𝟏}^2i\sigma _x^2).$$ (156) Then, after the first application of $`U`$, the system state becomes: $$U|00=\frac{1}{\sqrt{2}}(|00+i|11).$$ (157) Now let’s first consider some traditional moves of Alice and Bob, either cooperate (apply matrix $`U_A=U_B=\mathrm{𝟏}`$) or defect (apply the spin-flip Pauli matrix $`U_A=U_B=\sigma _x`$): $`\text{both cooperate: }(\mathrm{𝟏}\mathrm{𝟏})U|00={\displaystyle \frac{1}{\sqrt{2}}}(|00+i|11)`$ (158) $`\text{Alice defects: }(\sigma _x\mathrm{𝟏})U|00={\displaystyle \frac{1}{\sqrt{2}}}(|10+i|01)`$ (159) $`\text{Bob defects: }(\mathrm{𝟏}\sigma _x)U|00={\displaystyle \frac{1}{\sqrt{2}}}(|01+i|10)`$ (160) $`\text{both defect: }(\sigma _x\sigma _x)U|00={\displaystyle \frac{1}{\sqrt{2}}}(|11+i|00).`$ (161) Then when we apply the inverse of the unitary transformation $`U`$, namely $`U^1=U^{}`$, we get $`\text{both cooperate: }U^{}{\displaystyle \frac{1}{\sqrt{2}}}(|00+i|11)=|00\text{ with probability }1`$ (162) $`\text{Alice defects: }U^{}{\displaystyle \frac{1}{\sqrt{2}}}(|10+i|01)=|10\text{ with probability }1`$ (163) $`\text{Bob defects: }U^{}{\displaystyle \frac{1}{\sqrt{2}}}(|01+i|10)=|01\text{ with probability }1`$ (164) $`\text{both defect: }U^{}{\displaystyle \frac{1}{\sqrt{2}}}(|11+i|00)=|11\text{ with probability }1.`$ (165) These correspond to the four classical outcomes in Table VI, demonstrating that the classical game is encompassed by the quantum prisoner’s dilemma. Now let’s consider some less traditional quantum moves by Alice and Bob. For example, suppose Alice plays $`\mathrm{𝟏}`$ and Bob plays the Hadamard matrix $`H`$: $$(\mathrm{𝟏}H)U|00=\frac{1}{2}|0(|0+|1)+\frac{i}{2}|1(|0|1)=\frac{1}{2}[|00+|01+i|10i|11].$$ (166) Then applying $`U^{}`$ to the last equation we get the final state as $$U^{}(\mathrm{𝟏}H)U|00=\frac{1}{\sqrt{2}}(|01i|11).$$ (167) Since $`|\frac{1}{\sqrt{2}}|^2=\frac{1}{2}`$ and $`|\frac{i}{\sqrt{2}}|^2=\frac{1}{2}`$, a measurement of the latter state will give Alice a payout of 0 or a payout of 1 with equal probability, so $`\overline{\pi }_A=0.5`$, $`\overline{\pi }_B=3`$. Conversely, suppose Bob plays $`\mathrm{𝟏}`$ and Alice plays the Hadamard matrix $`H`$: $$(H\mathrm{𝟏})U|00=\frac{1}{2}[|00+|10+i|01i|11].$$ (168) Then applying $`U^{}`$ to the last equation we get the final state of the reversed play as $$U^{}(H\mathrm{𝟏})U|00=\frac{1}{\sqrt{2}}(|10i|11).$$ (169) A measurement of the latter state will give Alice a payout of 5 or a payout of 1 with equal probability, so $`\overline{\pi }_A=3`$, $`\overline{\pi }_B=0.5`$. We will summarize the remaining cases we want to consider: $`(H\sigma _x)U|00={\displaystyle \frac{1}{2}}[|01+|11+i|00i|10]`$ (170) $`(\sigma _xH)U|00={\displaystyle \frac{1}{2}}[|10+|11+i|00i|01]`$ (171) $`(HH)U|00={\displaystyle \frac{1}{\sqrt{2^3}}}[|00+|10+|01+|11+i|00i|10i|01+i|11],`$ (172) $`U^{}(H\sigma _x)U|00={\displaystyle \frac{1}{\sqrt{2}}}[|11i|10],\overline{\pi }_A=3,\overline{\pi }_B=0.5`$ (173) $`U^{}(\sigma _xH)U|00={\displaystyle \frac{1}{\sqrt{2}}}[|11i|01],\overline{\pi }_A=0.5,\overline{\pi }_B=3`$ (174) $`U^{}(HH)U|00={\displaystyle \frac{1}{2}}[|00+|11i|01i|10],\overline{\pi }_A=\overline{\pi }_B=2.25.`$ (175) Let ‘$``$’ denote ‘is preferred to’. Alice no longer has a preferred strategy. While $`\sigma _x_A\mathrm{𝟏}`$, if Bob plays $`\sigma _x`$ or $`H`$, then $`H_A\sigma _x`$. This is shown in Table VII. In addition, The payoff state $`(1,1)`$ corresponding to $`(\sigma _x,\sigma _x)`$ is no longer a Nash equilibrium. However, the outcome $`(2\frac{1}{4},2\frac{1}{4})`$ corresponding to $`(H,H)`$ is now a Nash equilibrium, although it is not Pareto optimal. Clearly the addition of quantum moves changes the game outcome. To induce Pareto optimality, let’s expand the set of allowed moves to be members of $`S=\{\mathrm{𝟏},\sigma _x,H,\sigma _z\}`$. The result is shown in Table VIII. The outcome $`(2\frac{1}{4},2\frac{1}{4})`$ is no longer a Nash equilibrium, but we have a new Nash equilibrium at $`(3,3)`$ corresponding to $`(\sigma _z,\sigma _z)`$. The payoffs are equal to those of the non-equilibrium strategy point $`(\mathrm{𝟏},\mathrm{𝟏})`$, so it is not jointly dominated. This Nash equilibrium is Pareto optimal. End of Prisoner’s Dilemma. What is the meaning of the unitary matrix $`U`$ that is applied at the beginning and end of the game? That remains to be determined. Sometimes it is ascribed to a third player, a referee or a co-ordinator. But there are other interpretations. Perhaps the best is that ‘it acts as a collaborator to the players and serves to maximize the payoff at the Nash equilibria’ CT . An Invisible Hand in prisoner’s dilemma? More work is needed. ## .13 Battle of the sexes game: a quantum game with entanglement The so-called ‘battle of the sexes’ game is not really a battle: it’s a love fest with conflicting values. Alice and Bob want to spend an evening together, and if they spend it apart, their respective payoffs are $`\{\gamma ,\gamma \}`$. As usual, Alice’s payoff is listed first and Bob’s payoff second. Alice prefers to spend the evening at the Opera (O), while Bob prefers to spend the evening watching TV (T). The payoffs for both at the Opera are $`\{\alpha ,\beta \}`$, while for both watching TV, the payoffs are $`\{\beta ,\alpha \}`$. It is assumed $`\alpha >\beta >\gamma `$. Alice and Bob are both at work at their respective jobs, and are not able to communicate (no cellphones). Each plans to show up either at the Opera or at Bob’s house for TV, in hopes of meeting the other at that place. The moves for each are thus members of the set $`\{O,T\}`$. The game is shown in Table IX. Inspection of the Table shows two Nash equilibria in moves: $`(O,O)`$ and $`(T,T)`$. A unilateral departure of either player from one of these equilibria results in a smaller payoff. However $`\mathrm{}`$, there is a Nash equilibrium in each row for Alice, and in each column for Bob. So how does either player decide what to do? In addition, there is a third hidden Nash equilibrium in mixed strategies resulting from Alice playing $`O`$ with probability $`p`$ and $`T`$ with probability $`1p`$, while Bob plays $`O`$ with probability $`q`$ and $`T`$ with probability $`1q`$, where $`p`$ and $`q`$ are neither $`0`$ nor $`1`$. Calculation shows $`p=\frac{\alpha \gamma }{\alpha +\beta 2\gamma }`$, while $`q=\frac{\beta \gamma }{\alpha +\beta 2\gamma }`$. These probabilities give the expected payoffs to Alice and Bob as $$\overline{\pi }_A(p,q)=\overline{\pi }_B(p,q)=\frac{\alpha \beta \gamma ^2}{\alpha +\beta 2\gamma }.$$ (176) In the corner Nash equilibria shown in Table IX, one of Alice or Bob receives a payoff of $`\alpha `$ and the other a payoff of $`\beta `$. But $`\alpha >\beta >\overline{\pi }_A(p,q)`$. So both Alice and Bob are worse off in the third Nash equilibrium. To find this third Nash equilibrium, we first write Alice’s expected payoff given the assumed probabilities of each move of Alice and Bob: $$\overline{\pi }_A=pq\alpha +p(1q)\gamma +(1p)q\gamma +(1p)(1q)\beta .$$ (177) Then, maximizing over $`p`$, $$\frac{\overline{\pi }_A}{p}=q\alpha +(1q)\gamma q\gamma (1q)\beta =0.$$ (178) Solving the latter equation for $`q`$ results in $`q=\frac{\beta \gamma }{\alpha +\beta 2\gamma }`$. A similar calculation maximizing Bob’s expected payoff yields $`p`$. How do quantum strategies change things? Let’s map $`|O|0`$ and $`|T|1`$,and then entangle states by applying our unitary matrix $`U`$, $$U=\frac{1}{\sqrt{2}}(\mathrm{𝟏}^2+i\sigma _x^2),$$ (179) to an initial state $`|00`$. Then, after the first application of $`U`$, the system state becomes: $$U|00=\frac{1}{\sqrt{2}}(|00+i|11),$$ (180) as before. Both Alice and Bob know $`U`$ and the initial state $`|00`$. We again allow Alice and Bob to make moves from the strategy set $`S=\{\mathrm{𝟏},\sigma _x,H,\sigma _z\}`$ on their individual qubits. And then we apply $`U^{}`$ to the result. The final states are those calculated previously in Prisoner’s Dilemma, but the expected payoffs are different, as shown in the following Table X. The upper left-hand entries show the classical game is contained in the quantum game. The only Nash equilibrium in the Table is $`(\beta ,\alpha )`$ corresponding to $`(\sigma _x,\sigma _x)`$. Alice and Bob spend an evening watching television together, with Alice having a payoff of $`\beta `$ less than Bob’s payoff of $`\alpha `$. At $`(\sigma _x,\sigma _x)`$ neither Alice nor Bob can unilaterally increase his or her payoff, and since this set of payoffs is not jointly dominated by another set of payoffs, it is also Pareto optimal. Television rules! It remains to consider mixed strategies. It is clear the four corner payoffs in the Table are the extreme points of a convex set. So we only need consider consider convex combinations of $`\mathrm{𝟏}`$ and $`\sigma _z`$. Alice’s expected payoff takes the form $$\overline{\pi }_A=pq\alpha +p(1q)\beta +(1p)q\beta +(1p)(1q)\alpha .$$ (181) Maximizing over $`p`$, $$\frac{\overline{\pi }_A}{p}=q\alpha +(1q)\beta q\beta (1q)\alpha =0.$$ (182) Solving for $`q`$ gives $`q=\frac{1}{2}`$. Similarly, $`p=\frac{1}{2}`$. The mixed strategies $`(\frac{1}{2}\mathrm{𝟏}+\frac{1}{2}\sigma _z,\frac{1}{2}\mathrm{𝟏}+\frac{1}{2}\sigma _z)`$ yield payoffs of $`(\frac{\alpha +\beta }{2},\frac{\alpha +\beta }{2})`$. At last equality between Bob and Alice! This Nash equilibrium is also Pareto optimal, as it is not jointly dominated by either $`(\alpha ,\beta )`$ or $`(\beta ,\alpha )`$. ## .14 Newcomb’s Game: a game against a Superior Being Alice plays the following game against a Superior Being (SB). The SB may be thought of as God, a superior intelligence from another planet, or as a supercomputer that is very good at predicting Alice’s thought processes SJB . There are two boxes $`B_1`$ and $`B_2`$. $`B_1`$ contains $1000. $`B_2`$ contains either $1,000,000 or $0, depending on which amount SB put in the box. Alice may choose to take either both boxes or only $`B_2`$. If the SB has predicted that Alice will choose both boxes, then SB puts $0 in $`B_2`$, while if the SB has predicted Alice will take only box $`B_2`$, then SB puts $1,000,000 in $`B_2`$. The game is depicted in Table XI. Alice clearly has a dominant strategy, which is to take both boxes, as each payoff in the second row is greater than the corresponding payoff in the first row. On the other hand, the dominant strategy conflicts with expected utility theory (here utility is taken to be linear in the payoffs). Suppose the predictive accuracy of SB is $`p`$. Then according to expected ultility theory, Alice will be indifferent between taking both boxes or only $`B_2`$ if $$p\text{ }\$1,000,000+(1p)\text{ }\$0=(1p)\text{ }\$1,001,000+p\text{ }\$1000.$$ (183) For $`p>.5005`$ Alice would prefer the strategy of only taking box $`B_2`$, conflicting with the dominant strategy. There are various ways to resolve this dilemma SJB . For example, if SB is omniscient (p=1), then the Table has only two entries, $1000 and $1,000,000. So automaton Alice will choose whichever SB has predicted, and the paradox is resolved. But here we are interested in the quantum game PS2 . SB surely knows the universe is based on quantum physics, not on classical physics, which is only the biased view of beings who are approximately two meters high. The quantum Newcomb’s game takes place in the Hilbert space $`𝐇_1𝐇_2`$, which we will take to be a 2-qubit space, with the left qubit denoting Alice’s actions, and the right qubit denoting the actions of the SB. For SB, $`|0`$ represents the placement of $1,000,000 in box $`B_2`$, while $`|1`$ represents the placement of $0 in $`B_2`$. For Alice, $`|0`$ represents taking $`B_2`$ only, while $`|1`$ represents taking both boxes. The basis vectors of $`𝐇_1𝐇_2`$ are $`|00`$, $`|01`$, $`|10`$, $`|11`$, corresponding to the payoff states in Table XI. The initial state of the game is $`\mathrm{\Lambda }=|00`$ if SB puts $1,000,000 in box $`B_2`$, or $`\mathrm{\Lambda }=|11`$ if SB puts nothing in $`B_2`$. The course of the game is as follow. *Step 1:* SB makes its choice, $`|0`$ or $`|1`$. Once made this choice cannot be altered. *Step 2:* SB applies the Hadamard matrix $`H`$ to Alice’s qubit; that is, the operator $`H\mathrm{𝟏}`$ to the initial state $`\mathrm{\Lambda }`$. *Step 3:* Alice applies the spin flip operator $`\sigma _x\mathrm{𝟏}`$ with probability $`w`$ or the identity matrix $`\mathrm{𝟏}\mathrm{𝟏}`$ with probability $`1w`$ to the current state of the game. (These operate only on her own qubit.) *Step 4:* The SB applies $`H\mathrm{𝟏}`$ to the current state of the game, and the payoff to Alice is determined. If the SB has chosen $`|0`$, then the sequence of steps in the game is as follow: $`(H\mathrm{𝟏})|00{\displaystyle \frac{1}{\sqrt{2}}}(|00+|10)`$ (184) $`w(\sigma _x\mathrm{𝟏})(H\mathrm{𝟏})|00{\displaystyle \frac{w}{\sqrt{2}}}(|00+|10)`$ (185) $`(w(\sigma _x\mathrm{𝟏})+(1w)(\mathrm{𝟏}\mathrm{𝟏}))(H\mathrm{𝟏})|00{\displaystyle \frac{1}{\sqrt{2}}}(|00+|10)`$ (186) $`(H\mathrm{𝟏})(w(\sigma _x\mathrm{𝟏})+(1w)(\mathrm{𝟏}\mathrm{𝟏}))(H\mathrm{𝟏})|00|00.`$ (187) Thus Alice takes only box $`B_0`$ and receives $1,000,000. The SB has correctly predicted Alice’s move. If the SB has chosen $`|1`$, then the sequence of steps in the game is as follow: $`(H\mathrm{𝟏})|11{\displaystyle \frac{1}{\sqrt{2}}}(|01|11)`$ (188) $`w(\sigma _x\mathrm{𝟏})(H\mathrm{𝟏})|11{\displaystyle \frac{w}{\sqrt{2}}}(|11|01)`$ (189) $`(w(\sigma _x\mathrm{𝟏})+(1w)(\mathrm{𝟏}\mathrm{𝟏}))(H\mathrm{𝟏})|11{\displaystyle \frac{12w}{\sqrt{2}}}(|01|11)`$ (190) $`(H\mathrm{𝟏})(w(\sigma _x\mathrm{𝟏})+(1w)(\mathrm{𝟏}\mathrm{𝟏}))(H\mathrm{𝟏})|11(12w)|11.`$ (191) The final value is maximized when $`w=0`$. Thus Alice takes both boxes and receives $1,000. The SB has again perfectly predicted Alice’s move. The SB did not require omiscience to achieve this result, only a knowledge of quantum mechanics. By applying the Hadamard matrix (the quantum Fourier transform) to the initial state of the game, the SB induced Alice to behave in a way so as to confirm the SB’s prediction. ## .15 Evolutionarily stable strategy game It seems that quantum games are played about us every day at a molecular level. Gogonea and Merz GMerz indicate games are being played at the quantum mechanical level in protein folding. Turner and Chao TC studied the evolution of competitive interactions among viruses in an RNA phage, and found the fitness of the phage generates a payoff matrix conforming to the two-person prisoner’s dilemma game. We want to briefly touch on some game theory aspects of biology. The concept of *evolutionarily stable strategy* (ESS), which we previously defined in connection with the concept of Nash equilibrium, was introduced into game theory EGTSE to deal with some problems in population biology and with the fact there may be multiple Nash equilibria. In *Evolution and the Theory of Games* MSETG Maynard Smith noted that ‘game theory is more readily applied to biology than to the field of economic behaviour for which it was originally designed’. Consider a population of $`N`$ members who are randomly matched in pairs to play a symmetric bimatrix (i.e., $`2\times 2`$) game. By *symmetric* is meant the following. Let $`S`$ be the set of player moves, and let $`s_i`$, $`s_j`$ be moves that are available to both Alice and Bob. Then Alice’s expected payoff when she plays $`s_i`$ and Bob plays $`s_j`$ is the same as Bob’s expected payoff if he plays $`s_i`$ and Alice plays $`s_j`$: $$\overline{\pi }_A(s_i,s_j)=\overline{\pi }_B(s_j,s_i).$$ (192) That is, Alice’s payoff matrix $`\mathrm{\Pi }_A`$ is the transpose of Bob’s payoff matrix: $`\mathrm{\Pi }_A=\mathrm{\Pi }_B^T`$. This defines the symmetry of the game. The game becomes *evolutionary* if over time moves $`s_i`$ with higher payoffs gradually replace those $`s_j`$ with lower payoffs. In such a game, Maynard Smith and Price MSP showed that a population which adopts an ESS can withstand a small invading group. But what if the current population, in equilibrium while playing classical moves, is invaded by a population playing quantum moves? This is the problem considered by Iqbal and Toor IT . Suppose the proportion of the population playing the move $`s_i`$ in a symmetric bimatrix game is $`p_i`$, while the proportion playing the move $`s_j`$ is $`p_j`$. Define the *fitness* $`w`$ of moves $`s_i`$ and $`s_j`$ as follows: $`w(s_i)=p_i\overline{\pi }(s_i,s_i)+p_j\overline{\pi }(s_i,s_j)`$ (193) $`w(s_j)=p_i\overline{\pi }(s_j,s_i)+p_j\overline{\pi }(s_j,s_j).`$ (194) The first equation says the fitness of move $`s_i`$ is a weighted average of the payoff to playing $`s_i`$ against an opponent also playing $`s_i`$ and of the payoff to playing $`s_i`$ against an opponent playing $`s_j`$. The respective weights are the proportions of the population playing $`s_i`$ and $`s_j`$. The second equation is really the same as the first with indexes switched. For our *quantum evolutionarily stable strategy game* we will assume that the symmetric bimatrix game played between the two population groups is the Prisoner’s Dilemma game. The payoff matrix for this game is that previously given in Table VI. Note that the payoff matrix of one player is the transpose of the payoff matrix of the other player, which is required for symmetry. Note also that the unitary matrix $`U=\frac{1}{\sqrt{2}}(\mathrm{𝟏}^2+i\sigma _x^2)`$ used in the quantum Prisoner’s Dilemma game is also symmetric between the two players. For classical moves, the payoff state $`\{s_A,s_B\}=\{D,D\}`$ and $`\{\pi (s_A),\pi (s_B)\}=\{1,1\}`$, which is a Nash equilibrium, is also an evolutionarily stable strategy. Consider, however, the effect of an invading force of mutants playing quantum moves. For ease of reference, we will reproduce Table VIII here as Table XII. We will label $`\{\mathrm{𝟏},\sigma _x\}`$ as classical moves, and $`\{H,\sigma _z\}`$ as mutant moves. We see that $`\sigma _x`$ is not evolutionarily stable against $`H`$. Members playing $`\sigma _x`$ will die out and the population will soon be comprised of mutants playing $`H`$. The new ESS will yield the payoff 2$`\frac{1}{4}`$ to either mutant party. If this new population is now invaded by different mutants playing $`\sigma _z`$, then $`H`$ is no longer an ESS. Members playing $`H`$ will die out, and the population will soon be comprised of mutants playing $`\sigma _z`$. These mutants will enjoy a payoff of 3, and will appear fat and happy when contrasted with the original population. ## .16 Card game: a quantum game without entanglement The following game doesn’t use entanglement, but is heuristic for its mathematical setup, and is good preparation for more complicated games that follow. Bob and Alice play the following card game JXHMXR . There are three cards, otherwise identical, except for the following markings: the first card has a circle on each side; the second card has a dot on each side; the third card has a circle on one side and a dot on the other. Alice puts the three cards in a black box and shakes it to randomize the three cards. Bob is allowed to blindly draw one card from the box. If it has the same mark on each side, Alice wins $`+1`$ from Bob. If the card has different marks on each side, Bob wins $`+1`$ from Alice. Of course, two of the cards having the same mark on each side, Alice has expected payoff $`\overline{\pi }_A=\frac{2}{3}(1)+\frac{1}{3}(1)=\frac{1}{3}`$, while Bob has expected payoff $`\overline{\pi }_B=\frac{1}{3}(1)+\frac{2}{3}(1)=\frac{1}{3}`$. The game is unfair to Bob. One way to make the game fair, in a classical sense, would be to allow Bob to look in the black box and see the upper faces of the three cards before drawing one of them. Then if Bob saw two circles facing up among the three cards, he would randomly draw one of those two cards, while if he saw two dots facing up, he would radomly draw one of the latter two cards. Since one of the two cards with identical upside marks must have different markings on each side, this would give Bob an expected payoff $`\overline{\pi }_B=0`$. The game would now be fair. However, we are not going to let Bob do this. In fact, it’s a black box so that he *can’t* look inside, but he can stick his hand in and pull one card out. Instead, to create the quantum equivalent of looking at the upper faces of all three cards, we are going to 1) allow Bob to make a single *query* to the black box or qubit database $`|r`$; and 2), allow Bob to withdraw from the game once he sees the upper face of the card he draws. This setup is highly artificial, and it is doubtful we are even describing the same game, but this quantized version of the Card Game will allow us to make several heuristic points. To describe the quantum game setup, let the card state be $`|0`$ if the card has a circle up, and $`|1`$ if a card has a dot up. The three-card state can be written as $$|r=|r_0r_1r_2$$ (195) where $`r_k\{0,1\}`$. As part of Bob’s query, we will require the following unitary matrix $`U_k`$: $$U_k=\left(\begin{array}{cc}1& 0\\ 0& e^{i\pi r_k}\end{array}\right).$$ (196) Note that if $`r_k=0`$, then $`U_k=\mathrm{𝟏}`$, while if $`r_k=1`$, then $`U_k=\sigma _z`$. Now we apply the Hadamard matrix $`H`$ to $`U_k`$ to form $`HU_kH`$ and obtain: $$HU_kH=\frac{1}{2}\left(\begin{array}{cc}1& 1\\ 1& 1\end{array}\right)\left(\begin{array}{cc}1& 0\\ 0& e^{i\pi r_k}\end{array}\right)\left(\begin{array}{cc}1& 1\\ 1& 1\end{array}\right)=\frac{1}{2}\left(\begin{array}{cc}1+e^{i\pi r_k}& 1e^{i\pi r_k}\\ 1e^{i\pi r_k}& 1+e^{i\pi r_k}\end{array}\right).$$ (197) Thus, applying this transformation to the state $`|0`$, we get $$HU_kH|0=\frac{1}{2}\left(\begin{array}{cc}1+e^{i\pi r_k}& 1e^{i\pi r_k}\\ 1e^{i\pi r_k}& 1+e^{i\pi r_k}\end{array}\right)\left(\begin{array}{c}1\\ 0\end{array}\right)=\frac{1}{2}\left(\begin{array}{c}1+e^{i\pi r_k}\\ 1e^{i\pi r_k}\end{array}\right)=\frac{1+e^{i\pi r_k}}{2}|0+\frac{1e^{i\pi r_k}}{2}|1.$$ (198) Note that if $`r_k=0`$, $`HU_kH|0=|0`$, while if $`r_k=1`$, $`HU_kH|0=|1`$. Thus, $$HU_kH|0=|r_k.$$ (199) So now let’s assume that Bob has a query machine that depends on state $`|r`$ in the black box. The machine has three inputs and gives three outputs. To determine the upside marks of the three cards, Bob inputs $`|000`$ to obtain: $$(HU_kHHU_kHHU_kH)|000=|r_0r_1r_2.$$ (200) So after Bob’s query, he knows the upside marks of the three cards: either some element of the set $`S_0=\{\text{ 3-qubit permuations of }\{|0,|0,|1\}\}`$ or some element of the set $`S_1=\{\text{ 3-qubit permuations of }\{|0,|1,|1\}\}`$. If $`S_0`$ descibes the state of the black box, then Bob knows the winning card has a circle on the upside face. If $`S_1`$ describes the state of the black box, then Bob know the winning card has a dot on the upwise face. So now Bob draws his card, and gets to look at the upside face only. If the drawn card has a circle on the upside face, and the black box $`S_0`$, then Bob has an equal chance of winning. But if the black box $`S_1`$, then Bob refuses to play because he knows the drawn card is a losing card. A similar analysis applies when the drawn card has a dot on the upside face. So a query to the database shows Bob whether there are two circles or two dots showing face up in the black box, and thus when he draws his card he knows that if it matches the two upside marks, then he has a 50-50 chance of winning, while if the drawn card doesn’t matched the two upside marks, the card is definitely a loser and he should exercise his option to withdraw from the game. With respect to entanglement, the operators $`H`$ and $`U_k`$ form simple linear combinations of qubits, while the quantum query machine is a tensor product of these operations. Hence there is no entanglement of states in this game. Du *et. al.* note that that the general rule appears to be that entanglement is required in static quantum games to make a difference from classical outcomes, but not in dynamic games. The key is the ability of the player to affect the state of others’ qubits. This can be done through entanglement or through the time steps of a dynamic game. ## .17 Quantum teleportation and pseudo-telepathy Alice and Bob are seven light-years apart and share an entangled pair of qubits, say $`|b_0=\frac{1}{\sqrt{2}}(|00+|11)`$. If Alice measures her qubit and finds it is in the state $`|0`$, then Bob’s qubit is guaranteed to be in the state $`|0`$ also. If Alice finds by measurement her qubit is in the state $`|1`$, then Bob’s qubit will also be found in the state $`|1`$. That is, *Alice’s measurement affects the state of Bob’s qubit*. As far as we know, this transmission of influence through the Bohr channel takes place instantaneously. It is not affected by distance or limited by the speed of light. It is spooky action at a distance. It is also the basis for quantum teleportation. $`\mathrm{𝐓𝐞𝐥𝐞𝐩𝐨𝐫𝐭𝐚𝐭𝐢𝐨𝐧}\text{ }`$. The quantum teleportation protocol BBCJPW , by contrast, does not take place instantaneously, since it uses a classical channel as well as a Bohr (EPR) channel. On the other hand, a quantum state disappears in one place and reappears in another: hence it is teleported. The traditional teleportation protocol works like this. Alice has an unknown quantum state $`|\psi `$ she wants to transmit to Bob. She will do this in two pieces: she will use an entangled Bohr channel, and an additional classical channel to transmit some classical bits. Alice and Bob have made previous arrangement to share an entangled pair of particles, this time say in the Bell state $`|b_3`$: $$|b_3=\frac{1}{\sqrt{2}}(|01|10).$$ (201) The unknown state Alice is trying to transmit may be written in terms of unknown amplitudes $`a`$, $`b`$, $`|a|^2+|b|^2=1`$, as $$|\psi =a|0+b|1.$$ (202) We may write the initial state of the 3-qubit system as: $`|\psi |b_3=(a|0+b|1)({\displaystyle \frac{1}{\sqrt{2}}}(|01|10))`$ (203) $`={\displaystyle \frac{a}{\sqrt{2}}}|001{\displaystyle \frac{a}{\sqrt{2}}}|010+{\displaystyle \frac{b}{\sqrt{2}}}|101{\displaystyle \frac{b}{\sqrt{2}}}|110.`$ (204) We want to rewrite this state in terms of the Bell basis, for reasons that will become apparent. To do this, we take the inner product of $`|\psi |b_3`$ with each of the Bell vectors in order to find the multiplier on each Bell state. Note that we take the inner product with the *two left-most* qubits in equation (184). These qubits are under the control of Alice. $`b_0|(|\psi |b_3)=+{\displaystyle \frac{a}{2}}|1{\displaystyle \frac{b}{2}}|0`$ (205) $`b_1|(|\psi |b_3)={\displaystyle \frac{a}{2}}|0+{\displaystyle \frac{b}{2}}|1`$ (206) $`b_2|(|\psi |b_3)=+{\displaystyle \frac{a}{2}}|1+{\displaystyle \frac{b}{2}}|0`$ (207) $`b_3|(|\psi |b_3)={\displaystyle \frac{a}{2}}|0{\displaystyle \frac{b}{2}}|1.`$ (208) Using these residual state multipliers, we can then write the state $`|\psi |b_3`$ in terms of the Bell basis: $$|\psi |b_3=\frac{1}{2}[\left(\begin{array}{c}b\\ +a\end{array}\right)|b_0+\left(\begin{array}{c}a\\ +b\end{array}\right)|b_1+\left(\begin{array}{c}+b\\ +a\end{array}\right)|b_2+\left(\begin{array}{c}a\\ b\end{array}\right)|b_3].$$ (209) Now let’s rewrite the last equation in terms of $`2\times 2`$ matrices: $`|\psi |b_3={\displaystyle \frac{1}{2}}[\left(\begin{array}{cc}0& 1\\ 1& 0\end{array}\right)\left(\begin{array}{c}a\\ b\end{array}\right)|b_0+\left(\begin{array}{cc}1& 0\\ 0& 1\end{array}\right)\left(\begin{array}{c}a\\ b\end{array}\right)|b_1+`$ (218) $`\left(\begin{array}{cc}0& 1\\ 1& 0\end{array}\right)\left(\begin{array}{c}a\\ b\end{array}\right)|b_2+\left(\begin{array}{cc}1& 0\\ 0& 1\end{array}\right)\left(\begin{array}{c}a\\ b\end{array}\right)|b_3].`$ (227) We can rewrite this again in terms of the Pauli spin matrices: $$|\psi |b_3=\frac{1}{2}[i\sigma _y\left(\begin{array}{c}a\\ b\end{array}\right)|b_0\sigma _z\left(\begin{array}{c}a\\ b\end{array}\right)|b_1+\sigma _x\left(\begin{array}{c}a\\ b\end{array}\right)|b_2\mathrm{𝟏}\left(\begin{array}{c}a\\ b\end{array}\right)|b_3].$$ (228) Now, to teleport her qubit to Bob, Alice must couple the unknown state $`|\psi `$ with her member of the entangled qubit pair. To do this she makes a joint (von Neumann) measurement of these two qubits, which comprise the two left-most qubits of $`|\psi |b_3`$. Alice’s measurement projects her two qubits into one of the four Bell states. This destroys the unknown state $`|\psi `$. But not to worry. Alice’s measurement also leaves Bob’s qubit in one of the following four states: $`|\psi |b_3|b_0\text{ Bob’s qubit }=i\sigma _y\left(\begin{array}{c}a\\ b\end{array}\right)`$ (231) $`|\psi |b_3|b_1\text{ Bob’s qubit }=\sigma _z\left(\begin{array}{c}a\\ b\end{array}\right)`$ (234) $`|\psi |b_3|b_2\text{ Bob’s qubit }=\sigma _x\left(\begin{array}{c}a\\ b\end{array}\right)`$ (237) $`|\psi |b_3|b_3\text{ Bob’s qubit }=\mathrm{𝟏}\left(\begin{array}{c}a\\ b\end{array}\right).`$ (240) Alice then, through a classical channel, transmits to Bob the results of her measurement: i.e., the Bell state she obtained. Then Bob applies the corresponding spin operator (which is its own inverse) to his qubit to recover the state $`|\psi =\left(\begin{array}{c}a\\ b\end{array}\right)`$: $`i\sigma _y`$ for $`|b_0`$, $`\sigma _z`$ for $`|b_1`$, $`\sigma _x`$ for $`|b_2`$, or $`\mathrm{𝟏}`$ for $`|b_3`$. (Actually, the overall signs \[signs that multiply both $`a`$ and $`b`$ equally\] don’t matter, since $`|\psi `$ is the same state as $`|\psi `$. So, for example, multiplication by $`\sigma _z`$ or by $`\mathrm{𝟏}`$ is sufficient.) To summarize, Alice and Bob share an entangled state $`|\theta `$ of two qubits. Alice wishes to teleport an unknown state $`|\psi `$ to Bob. To do this, she first performs a measurement of $`|\psi |\theta `$ in the Bell basis on her two qubits (the unknown state, and her qubit in the entangled state). She transmits the information of which Bell state she obtained to Bob. Bob applies the corresponding Pauli spin operator to his qubit and recovers the unknown state $`|\psi `$. $`\mathrm{𝐏𝐬𝐞𝐮𝐝𝐨}\mathrm{𝐭𝐞𝐥𝐞𝐩𝐚𝐭𝐡𝐲}\text{ }`$. ‘Entanglement is perhaps the most non-classical manifestation of quantum mechanics. Among its many interesting applications to information processing, it can be harnessed to *reduce* the amount of communication required to proces a variety of distributed computational tasks. Can it be used to *eliminate* communication altogether? Even though it cannot serve to signal information between remote parties, there are distributed tasks that can be performed without any need for communication, provided the parties share prior entanglement: this is the realm of *pseudo-telepathy*.’ BBT Consider the following *Pseudo-Telepathy Game* $`\mathrm{\Gamma }_N`$ between $`N`$ players. Since there are more than two players, we can’t call them Alice and Bob, so we’ll let them all be subscript Alices: $`A_1,A_2,\mathrm{},A_N`$. There are also two functions $`f`$ and $`g`$, each of which take $`N`$-qubit inputs. The game has the following steps. *Step 1*: The players mingle, discuss strategy, share random variables (in the classical setting) or entanglement (in the quantum setting). *Step 2*: The players separate and are not allowed to engage in any form of communication. Each player $`A_i`$ is given a single qubit input $`x_i`$ and requested to produce the single qubit output $`y_i`$. The players *win* $`+1`$ if $$f(x_1,x_2,\mathrm{},x_N)=g(y_1,y_2,\mathrm{},y_N).$$ (241) else they lose this amount. The functions $`f`$ and $`g`$ are defined as followings. Players are guaranteed that the sum of the qubits they are given is an even number: $`_ix_i`$ is even. (Think of what this means. If $`_ix_i`$ is even, then it is divisible by 2. Thus $`\frac{1}{2}_ix_i`$ is a whole number that is either odd or even. If odd, then $`\frac{1}{2}_ix_i\text{ mod }2=1`$. If even, then $`\frac{1}{2}_ix_i\text{ mod }2=0`$. But the latter case means $`\frac{1}{2}_ix_i\text{ mod }2`$ is also divisible by two, so that the original sum $`_ix_i`$ was divisible by 4.) The players are asked to produce an even sum of output bits $`_iy_i`$ if and only if the sum of the input bits $`_ix_i`$ is divisible by 4. Thus the criterion for the $`N`$-players to win is: $$\underset{i}{}y_i\text{ mod }2=\frac{1}{2}\underset{i}{}x_i\text{ mod }2.$$ (242) The left-hand side of this equation is $`g`$ and the right-hand side $`f`$. A win depends solely on the global state of the $`N`$ qubits, even though each player controls only $`1`$ qubit, and is not allowed to communicate with the other players. Note that the expected payoff to the players if any player $`i`$ randomizes the submission of $`y_i`$ is $`0`$, as mod $`2`$ produces only two outcomes. This is a very nice game, because it highlights the issue of cooperation between players, and because the game is scalable to any number $`N`$ of players. Now, the amazing thing is that if the players are allowed to share prior entanglement, as in Step 1, then they always win $`\mathrm{\Gamma }_N`$. To see how they do this, we need as components the Bell states $`|b_0`$ and $`|b_2`$, the Hadamard transform $`H`$, and the unitary or rotation matrix introduced in the Card Game, except here we will define it as: $$U_{\frac{\pi }{2}}=\left(\begin{array}{cc}1& 0\\ 0& e^{i\frac{\pi }{2}}\end{array}\right)=\left(\begin{array}{cc}1& 0\\ 0& i\end{array}\right),$$ (243) remembering that $`cos(\frac{\pi }{2})+i\text{ }sin(\frac{\pi }{2})=i`$. Note that $`U_{\frac{\pi }{2}}|0=|0`$ but $`U_{\frac{\pi }{2}}|1=i\text{ }|1`$. Since $`N`$ players share the entangled Bell states, the latter will have to be $`N`$-qubit Bell states. Let’s write our $`N`$-qubit Bell states in the following simplified form: $`|b_0^N={\displaystyle \frac{1}{\sqrt{2}}}(|0^N+|1^N)`$ (244) $`|b_2^N={\displaystyle \frac{1}{\sqrt{2}}}(|0^N|1^N).`$ (245) The first $`N`$-qubit state, $`|b_0^N`$ is the entangled state that all players agree to share. The second state may evolve in the course of play. Consider now the effect of the unitary matrix operating on a single qubit of $`|b_0^N`$: $$U_{\frac{\pi }{2}}|b_0^N=\frac{1}{\sqrt{2}}(|0^N+i\text{ }|1^N).$$ (246) The powers of $`i`$ are $`i,\text{ }i^2=1,\text{ }i^3=i,\text{ }i^4=1`$. So if $`U_{\frac{\pi }{2}}`$ is applied to two qubits, the sign on $`|1^N`$ becomes $`1`$, and thus $`|b_0^N|b_2^N`$. If applied to four qubits, the sign is unchanged, so $`|b_0^N|b_0^N`$. So if $`m`$ players apply $`U_{\frac{\pi }{2}}`$ to their individual qubits, the initial state $`|b_0^N`$ will remain unchanged if $`m=0\text{ mod }4`$. If $`m=2\text{ mod }4`$, then $`|b_0^N|b_2^N`$. If each player applies the Hadamard matrix to his qubit when the entangled state is $`|b_0^N`$, the result is a superposition of all states *with an even number of 1 bits*: $$(H^N)|b_0^N=\frac{1}{\sqrt{2^{N1}}}\underset{even\text{ }bit\text{ }y}{\overset{2^N1}{}}|y.$$ (247) Note that this does *not* mean the states $`|y`$ in the summation are even numbers. For example, $`|101=|5`$ is an odd number, but has an even number of 1 bits, while $`|100=|4`$ is an even number, but has an odd number of 1 bits. To see that the $`N`$-fold Hadamard transform (the Walsh transform) turns Bell state $`|b_0^N`$ into a superposition of even-bit numbers (meaning an even number of 1 bits), consider Table XIII, which is an analog of Table V. Note that the minus signs appear on the numbers with an odd number of 1 bits. So if we apply $`(HHH)`$ to $`\frac{1}{\sqrt{2}}(|000+|111)`$, we get $`\frac{1}{\sqrt{2^4}}(|0+|1+|2+|3+|4+|5+|6+|7+|0|1|2+|3|4+|5+|6|7)=\frac{2}{\sqrt{2^4}}(|0+|3+|5+|6`$), a superposition of numbers all of which have an even number of 1 bits. If the state has evolved to the state $`|b_2^N`$ due to player action, and each player applies the Hadamard matrix to his qubit, then the result is a superposition of all odd bit states (meaning states with an odd number of 1 bits): $$(H^N)|b_2^N=\frac{1}{\sqrt{2^{N1}}}\underset{odd\text{ }bit\text{ }y}{\overset{2^N1}{}}|y.$$ (248) So here, then, are the steps each player takes with respect to his or her qubit in the game $`\mathrm{\Gamma }_N`$: *Player Step 2a:* If a player receives qubit $`x_i=1`$, the player applies $`U_{\frac{\pi }{2}}`$ to his or her qubit in the entangled Bell state $`|b_0^N`$. Otherwise the player does nothing. *Consequence:* Because the sum of bits $`_ix_i`$ is even, an even number of players will perform this step. If $`_ix_i`$ is divisible by 4, then the Bell state $`|b_0^N`$ is left unchanged. But if $`_ix_i=2\text{ mod }4`$ then $`|b_0^N|b_2^N`$. *Player Step 2b:* Each player applies the Hadamard matrix $`H`$ to his or her qubit. *Consequence:* If the entangled state is still in the state $`|b_0^N`$ from Step 2a, then this present step transforms the entangled state into a superposition of all *even* bit states. But if the entangled state has been transformed into $`|b_2^N`$, then this step transforms the entangled state into a superposition of all *odd* bit states. *Player Step 2c:* Each player now measures his qubit in the computational basis ($`|0`$ vs. $`|1`$) to produce $`y_i`$. If $`_ix_i`$ was divisible by 4, the entangled qubit is in a superposition of even bit states, so will be projected under the measurement into a number with an even number of 1 bits. The players win, because $`_iy_i\text{ mod }2=0`$. If $`_ix_i=2\text{ mod }4`$, then the entangled qubit is in a superposition of odd bit states, so will be projected under the measurement into a number with an odd number of 1 bits. The players win again, because $`_iy_i\text{ mod }2=1`$. The players have demonstrated pseudo-telepathy by acting as though each knew what the other was doing, even though there was no communication between players. This was made possible by the shared entangled state $`|b_0^N`$ acting as a quantum invisible hand. We may characterize this pseudo-telepathy game in terms of traditional $`N`$-person game theory as follows. No player can secure any value by himself, so the value of a one-person coalition $`\{i\}`$ is $`0`$: $`v\{i\}=0`$. The value of the coalition of all players is $`1`$: $`v(N)=1`$. Such a game is said to be in $`(0,1)`$-*normalization*. Let $`S`$ be a subset of the set of players $`N`$. If for all $`SN`$ either $`v(S)=0`$ or $`v(S)=1`$, a game is said to be *simple*. Thus the pseudo-telepathy game is also simple; indeed $`v(S)=0`$ for all $`S`$ save $`S=N`$. Finally, a game is said to be constant sum if $`v(S)+v(NS)=v(N)`$. The pseudo-telepathy game is *not* constant sum, as $`v(S)+v(NS)=0`$ for $`SN`$, but $`v(N)=1`$. The set of imputations for this game is the set of probability vectors $`P=\{p_1,p_2,\mathrm{},p_N\}`$. This fulfills the requirement that $`_{iN}p_i=v(N)=1`$, and also the requirement that $`p_iv(\{i\})=0`$, for all $`iN`$. None of these allocation vectors is dominated by another, for $`SN`$. Thus the *core* of this game is the convex set of probability vectors $`P`$. ## .18 Quantum secret sharing The IRA has some secret information they want to preserve among their members, but are fearful that some of them may be MI5 informants, and that others may be arrested and reveal what they know under interrogation. So they need a secure way to embed the secret among themselves. A $`(k,n)`$ *threshold* scheme CGL is one in which any $`kn`$ members can reconstruct a secret, but $`k1`$ members cannot find *any* information about the secret at all. Let’s first, however, consider a simple example where two parties must cooperate to discover a secret quantum state HBB . Alice, Bob, and Gerald share the following entangled state (the left qubit is Alice’s, the right qubit is Gerald’s): $$|\psi =\frac{1}{\sqrt{2}}(|000+|111).$$ (249) First note we can rewrite this in terms of a different basis. Let $`|x^+={\displaystyle \frac{1}{\sqrt{2}}}(|0+|1)`$ (250) $`|x^{}={\displaystyle \frac{1}{\sqrt{2}}}(|0|1).`$ (251) This implies the reciprocal relations $`|0={\displaystyle \frac{1}{\sqrt{2}}}(|x^++|x^{})`$ (252) $`|1={\displaystyle \frac{1}{\sqrt{2}}}(|x^+|x^{}).`$ (253) So the original state in terms of the new basis would be $$|\psi =\frac{1}{2\sqrt{2}}[(|x^+x^++|x^{}x^{})(|0+|1)+(|x^+x^{}+|x^{}x^+)(|0|1)].$$ (254) Alice wishes to send a secret qubit $`|\varphi _{secret}=a|0+b|1`$ to Bob and Gerald in such a way that Bob and Gerald must cooperate in order to learn the secret. She essentially does this through the teleportation protocol, but we will also need the definitions of $`(|x^+,|x^{})`$ for part of the procedure. Alice combines the secret qubit $`|\varphi _{secret}`$ with the shared state $`|\psi `$ to form the overall state $$|\varphi _{secret}|\psi =\frac{1}{\sqrt{2}}(a|0000+b|1000+a|0111+b|1111).$$ (255) Alice now rewrites this in terms of the Bell basis. The multipliers on the Bell states are: $`b_0|(|\varphi _{secret}|\psi )={\displaystyle \frac{a}{2}}|00+{\displaystyle \frac{b}{2}}|11`$ (256) $`b_1|(|\varphi _{secret}|\psi )={\displaystyle \frac{a}{2}}|11+{\displaystyle \frac{b}{2}}|00`$ (257) $`b_2|(|\varphi _{secret}|\psi )={\displaystyle \frac{a}{2}}|00{\displaystyle \frac{b}{2}}|11`$ (258) $`b_3|(|\varphi _{secret}|\psi )={\displaystyle \frac{a}{2}}|11{\displaystyle \frac{b}{2}}|00.`$ (259) Alice now measures her two qubits in the Bell basis, sends the result to Gerald, and tells Bob to measure his qubit in the $`(|x^+,|x^{})`$ basis. After Alice’s Bell measurement, the qubits of Bob and Gerald will be in one of the following states: $`|b_0a|00+b|11`$ (260) $`|b_1a|11+b|00`$ (261) $`|b_2a|00b|11`$ (262) $`|b_3a|11b|00.`$ (263) If Bob gets $`|x^+`$ upon his measurement, then Gerald’s qubit becomes $`a|00+b|11a|0+b|1`$ (264) $`a|11+b|00a|1+b|0`$ (265) $`a|00b|11a|0b|1`$ (266) $`a|11b|00a|1b|0`$ (267) while if Bob gets $`|x^{}`$, Gerard’s qubit becomes $`a|00+b|11a|0b|1`$ (268) $`a|11+b|00a|1+b|0`$ (269) $`a|00b|11a|0+b|1`$ (270) $`a|11b|00a|1b|0.`$ (271) To reconstruct Alice’s qubit, Gerald needs to know what measurement Bob obtained, so that Gerald can apply the appropriate Paul spin matrix to his final qubit state. Thus Gerald and Bob together can reconstruct Alice’s qubit, but neither can do so alone. The appropriate Pauli spin matrices to be applied to Gerald’s final state are: Now that we have seen the close relation of quantum secret sharing to teleportation, at least in one example, let’s return to the $`(k,n)`$ threshold notion, and consider an example of a $`(2,3)`$ threshold scheme. This scheme works by splitting up a state among three parties in such a way that any two can reconstruct the original state. We begin with an unknown secret state that is not a qubit, but rather a *qutrit*. A qutrit is a ternary ‘trit’ that can take values in the three-dimensional Hilbert space spanned by $`(|0,|1,|2)`$. We’ve simply added one more dimension to a qubit. Note that for this example, tensor products expand by powers of 3, so 3 qutrits occupy a Hilbert space of dimension 27: $`𝐇_{\mathrm{𝟐𝟕}}=𝐇_\mathrm{𝟑}𝐇_\mathrm{𝟑}𝐇_\mathrm{𝟑}`$. We have an secret state $`|\varphi _{secret}=\alpha |0+\beta |1+\gamma |2`$. We have an encoding transformation that maps this 1-qutrit state into a mixed 3-qutrit state: $$|\varphi _{secret}\alpha (|000+|111+|222)+\beta (|012+|120+|201)+\gamma (|021+|102+|210).$$ (272) Now we can split this mixed 3-qutrit state between Alice, Bob, and Gerald. The left qutrit belongs to Alice, and the right qutrit to Gerald. Given their qutrits, no one has any idea about the original state, because the state they posses has an equal mixture of $`|0`$, $`|1`$, and $`|2`$. However, any two people can reconstruct the secret state $`|\varphi _{secret}`$. For example, Alice and Bob get together. Alice adds her qutrit to Bob’s modulo 3, then Bob adds his (new) qutrit to Alice’s. The result is the state $$(\alpha |0+\beta |1+\gamma |2)(|00+|12+|21).$$ (273) To see this, let’s consider just the multipliers on $`\alpha `$. When Alice and Bob get together, they have $$\alpha (|000+|111+|222)+\mathrm{}.$$ (274) Adding Alice’s qutrit to Bob’s modulo 3 we get $$\alpha (|000+|111+|222)+\mathrm{}\alpha (|000+|121+|212)+\mathrm{}.$$ (275) Then adding Bob’s (new) qutrit to Alice’s we get $`\alpha (|000+|121+|212)+\mathrm{}\alpha (|000+|021+|012)+\mathrm{}`$ (276) $`=(\alpha |0+\mathrm{})(|00+|12+|21).`$ (277) Alice’s qutrit is now identical with the secret state $`|\varphi _{secret}`$, which has been disentangled from the other qutrits. By a similar process Gerald and Bob could recover the secret state, or Alice and Gerald. ## .19 The density matrix and quantum state estimation The ‘No Cloning Theorem’ forbids a quantum copier of the following sort: the copier takes one quantum state as input and outputs two systems of the same kind. The no cloning theorem got its name after Nick Herbert proposed a faster-than-light communication device, published in *Foundations of Physics* in 1982 NH82 . This generated widespread attention and a flaw in the argument was soon found: the device required quantum cloning, and there were problems with producing identical copies of a quantum state. (Further background is found in AP2004 .) However, that is not the whole story. Preparing virtually identical copies is no problem, if we don’t try to do it in a single measurement. By statistical procedures the input state can be determined to any degree of accuracy. For example, for the unknown state $`|\psi `$, $$|\psi =a|0+b|1$$ (278) repeated measurement of $`n`$ such prepared states in the computational basis will yield $`|0`$ $`n_a`$ times and $`|1`$ $`n_b`$ times, where $`n_a+n_b=n`$. Then clearly $`{\displaystyle \frac{n_a}{n}}|a|^2=|\psi |0|^2`$ (279) $`{\displaystyle \frac{n_b}{n}}|b|^2=|\psi |1|^2.`$ (280) That is, the $`n`$ measurements will yield $`(x_1,x_2,\mathrm{},x_n)`$, where each $`x_i`$ is either $`0`$ or $`1`$. This corresponds to a set of Bernoulli trials whose Likelihood Function is $$L(p)=\underset{i=1}{\overset{n}{}}p^{x_i}q^{1x_i}=p^{{\scriptscriptstyle x_i}}q^{n{\scriptscriptstyle x_i}}.$$ (281) where p is the probability of $`1`$ and $`q=1p`$ is the probability of $`0`$. Maximizing $`L(p)`$ yields the estimate for $`p`$ as $$\widehat{p}=\frac{1}{n}x_i=\frac{n_b}{n}.$$ (282) This leads to the statistically-based *density matrix* $`\rho `$: $$\rho =\left(\begin{array}{cc}\frac{n_a}{n}& 0\\ 0& \frac{n_b}{n}\end{array}\right)=\frac{n_a}{n}\left(\begin{array}{cc}1& 0\\ 0& 0\end{array}\right)+\frac{n_b}{n}\left(\begin{array}{cc}0& 0\\ 0& 1\end{array}\right)=\frac{n_a}{n}|00|+\frac{n_b}{n}|11|.$$ (283) From the statistical point of view, the quantum state is a mathematical encoding of all data that can be collected this way. Before proceeding further we need to explain the differences between *pure* states and *mixed* states. If a quantum state $`|\psi `$ is a convex combination of other quantum states, it is said to be in a *mixed* state. Note that mixture involves classical probabilities or combinations, not amplitudes. But if a state $`|\psi `$ cannot be expressed as a convex combination of other states, it is said to be in a *pure* state. Pure states are the extreme points of a convex set of states. For a pure state $`|\varphi `$, the ket-bra $`|\varphi \varphi |`$ is called a *projection operator*. It projects $`|\varphi `$ onto itself ($`|\varphi \varphi |\varphi =|\varphi `$), and any state $`|\theta `$ orthogonal to $`|\varphi `$ is projected onto $`0`$ ($`|\varphi \varphi |\theta =0`$). For a pure state $`\varphi `$, the density matrix is simply $`\rho =|\varphi \varphi |`$. For a mixed state, where the system will be found in one of the extreme points $`|\varphi _j`$ with probability $`p_j`$, the *density matrix* $`\rho `$ is defined as the sum of the projectors weighted with the respective probabilities: $$\rho =\underset{j}{}p_j|\varphi _j\varphi _j|.$$ (284) Since the probabilities are non-negative and sum to one, this means $`\rho `$ is a positive semidefinite Hermitian operator (the eigenvalues are non-negative) and the trace of $`\rho `$ (the sum of the diagonal elements of the matrix, i.e. the sum of its eigenvalues) is equal to one. For example, let the pure state $`|\psi `$ be $`|\psi =a|0+b|1`$, where $`a`$ and $`b`$ are complex numbers with respective complex conjugates $`a^{}`$ and $`b^{}`$. Then the density matrix $`\rho `$ for $`|\psi `$ is $$\rho =|\psi \psi |=\left(\begin{array}{cc}aa& ab\\ ba& bb\end{array}\right).$$ (285) For $`a=\sqrt{\frac{2}{3}}`$, $`b=\sqrt{\frac{1}{3}}`$, this becomes $$\rho =|\psi \psi |=\left(\begin{array}{cc}\frac{2}{3}& \frac{\sqrt{2}}{3}\\ \frac{\sqrt{2}}{3}& \frac{1}{3}\end{array}\right).$$ (286) A measurement of $`|\psi `$ in the computational basis will yield $`|0`$ with probability $`\frac{2}{3}`$ or $`|1`$ with probability $`\frac{1}{3}`$. These probabilities are found in the trace of $`\rho `$. We may rewrite $`\rho `$ as $`\rho =\frac{2}{3}|00|+\frac{1}{3}|11|`$, losing any information in the off-diagonal elements. (This is what happens, as we shall see, during cloning.) Note that *after* the measurement, then either $`|\psi =|0`$ with probability 1, or $`|\psi =|1`$ with probability 1. As another example, suppose $`\frac{3}{4}`$ of the states in an ensemble of states are prepared in the state $`|\psi _1=.8|0+.6|1`$, while $`\frac{1}{4}`$ are prepared in the state $`|\psi _2=.6|0.8i|1`$. Then the density matrix for this mixed ensemble, using equation (240), is $$\rho =.75|\psi _1\psi _1|+.25|\psi _2\psi _2|=\left(\begin{array}{cc}.57& .36+12i\\ .3612i& .43\end{array}\right).$$ (287) A particle drawn from this ensemble and measured in the $`(|0,|1)`$ basis will be found in state $`|0`$ with probability .57 or in state $`|1`$ with probability .43. But if we wanted to use $`\rho `$ to find the probabilities for a *different* basis, we would need the off diagonal elements as well as the trace. To see this, suppose we draw a particle from the same ensemble and take a measurement in the orthonormal basis $`(|\varphi _1,|\varphi _2)`$, where $`|\varphi _1=.6|0+.8|1`$ and $`|\varphi _2=.8|0.6|1`$. Note that $`\varphi _1|\varphi _2=0`$ and $`|\varphi _1|\varphi _1|^2=|\varphi _2|\varphi _2|^2=1`$. Then $`\rho `$ gives as the probabilities $`P`$ of observing $`|\varphi _1`$ and $`|\varphi _2`$ as $`P(|\varphi _1)=(.6,\text{ }.8)\rho \left(\begin{array}{c}.6\\ .8\end{array}\right)=.826`$ (290) $`P(|\varphi _2)=(.8,\text{ }.6)\rho \left(\begin{array}{c}.8\\ .6\end{array}\right)=.174.`$ (293) Suppose we choose an observable $`\mathrm{}`$, such as the spin state of an electron. Then in the von Neumann formulation of quantum measurement, each observable is associated with a Hermitian operator $`A`$, with $`A|\psi _j=a_j|\psi _j`$, where $`|\psi _j`$ are the eigenvectors of $`A`$, and $`a_j`$ are the eigenvalues. Thus, using the same basis for $`\rho `$ and $`A`$, namely the eigenvectors of $`A`$, we have $$A\rho =\underset{j}{}p_jA|\psi _j\psi _j|=\underset{j}{}p_ja_j|\psi _j\psi _j|.$$ (294) Now the expected value of $`A`$, $`\overline{A}`$, is simply $$\overline{A}=\underset{j}{}p_j\text{ }a_j.$$ (295) Thus the latter may be represented as $$\overline{A}=\text{ trace }(A\rho ).$$ (296) There are many approaches to *quantum state estimation* via the density matrix $`\rho `$. The problem of state estimation is closely related to the problem of cloning, and is connected to issues of entanglement. The *maximum likelihood* approach considered earlier is probably the best. For the heuristic purposes of this essay a *Bayesian* framework MSred is revealing. We might start with the principle of indifference, or insufficient reason, and make the initial assumption that the density matrix has the fully mixed form (for a system in $`𝐇_2`$): $$\rho =\frac{1}{2}\mathrm{𝟏}=\left(\begin{array}{cc}\frac{1}{2}& 0\\ 0& \frac{1}{2}\end{array}\right).$$ (297) This corresponds to an ensemble, half of which are in an up state and half of which are in a down state: $$\rho =\frac{1}{2}|uu|+\frac{1}{2}|dd|=\frac{1}{2}\left(\begin{array}{c}1\\ 0\end{array}\right)(1\text{ }0)+\frac{1}{2}\left(\begin{array}{c}0\\ 1\end{array}\right)(0\text{ }1)=\frac{1}{2}\left(\begin{array}{cc}1& 0\\ 0& 0\end{array}\right)+\frac{1}{2}\left(\begin{array}{cc}0& 0\\ 0& 1\end{array}\right)=\frac{1}{2}\mathrm{𝟏}.$$ (298) Or we may start with the general form of the density matrix, which can be written in terms of the Pauli spin matrices and real numbers $`r_x`$, $`r_y`$, and $`r_z`$ as follows: $`\rho ={\displaystyle \frac{1}{2}}(\mathrm{𝟏}+𝐫\sigma )`$ (299) $`={\displaystyle \frac{1}{2}}(\mathrm{𝟏}+r_x\text{ }\sigma _x+r_y\text{ }\sigma _y+r_z\text{ }\sigma _z)`$ (300) $`={\displaystyle \frac{1}{2}}\left(\begin{array}{cc}1+r_z& r_xir_y\\ r_x+ir_y& 1r_z\end{array}\right).`$ (303) Here we require that the determinant of $`\rho `$ be non-negative, $`\mathrm{𝐝𝐞𝐭}\text{ }\rho 0`$, which implies $`\frac{1}{4}[1(r_x^2+r_y^2+r_z^2)]0`$, or that $`𝐫^2=r_x^2+r_y^2+r_z^21`$, so that each density matrix may be associated with a ball of radius $`1`$, called a *Bloch sphere*. Points on the surface of the ball correspond to pure states, while interior points correspond to mixed states. If we assume this form of the density matrix $`\rho `$ and then measure spin in the $`z`$ direction, obtaining a series of $`n`$ results $`u`$ and $`d`$ with frequencies $`n_u`$ and $`n_d`$, then the likelihood is $$L(n_u)=[\frac{1}{2}(1+r_z)^{\frac{n_u}{n}}][\frac{1}{2}(1r_z)^{\frac{nn_u}{n}}].$$ (304) Now consider the following *State Discrimination Game* $`\mathrm{\Gamma }_{sd}`$. There are $`N`$ states, members of the set $`S=\{|\psi _j,\text{ }j=0,1,\mathrm{},N1\}`$. Each of these states is represented by a density matrix $`\rho _j=\eta _j|\psi _j\psi _j|`$. Alice prepares a state $`\rho _k`$, unknown to Bob, and forwards it to Bob, along with the information that the associated $`|\psi _k`$ is a member of $`S`$. She also tells him the probabilities $`\eta _j`$ of each state in $`S`$. The $`\eta _j`$ are called *prior probabilities*. This, of course, immediately suggests a Bayesian framework, so let’s consider a Bayesian strategy called *quantum hypothesis testing* AC . Because there are $`N`$ states, Bob will follow a procedure that gives him $`N`$ outcomes, which we will label $`a_j`$. If Bob obtains outcome $`a_m`$ he will assume that the state he was sent was $`\rho _m`$. There is error probability $`p_E`$ that $`\rho _m\rho _k`$ and probability $`1p_E=p_D`$ that $`\rho _m=\rho _k`$. To complete the game description, we need to define the *channel matrix* $`[h(a_m|\rho _k)]`$ which expresses the probabilities that Bob will find $`a_m`$ given that $`\rho _k`$ was sent, and the *cost matrix* $`[c_{mk}]`$ which assigns a cost to making the hypothesis $`a_m`$ when $`\rho _k`$ was sent. No matter what $`\rho _k`$ was sent, Bob’s measurement will yield one of the $`a_m`$. This gives rise to the completeness condition that $$\underset{m=1}{\overset{N}{}}h(a_m|\rho _k)=1.$$ (305) Then the total error probability is $$p_E=1\underset{k=1}{\overset{N}{}}\eta _kh(a_k|\rho _k).$$ (306) The average amount $`c_B`$ Bob will pay Alice is given by the Bayesian cost matrix $$c_B=\underset{mk}{}\eta _kc_{mk}h(a_m|\rho _k).$$ (307) Bob’s goal is to minimize $`c_B`$. The only thing Bob controls are the elements in the channel matrix $`h`$. Thus Bob’s problem is $$min_{\text{ }𝐡}\text{ }\underset{mk}{}\eta _kc_{mk}h(a_m|\rho _k).$$ (308) This puts quantum state *discrimination* (finding a state in a given set of states) in the context of game theory. If we set the diagonal elements of the cost matrix equal to 0 (Bob pays nothing for being correct) and the other elements equal to a constant $`c`$ (all errors cost the same) then, comparing equations (256) and (257), Bob’s problem reduces to $$min_{\text{ }𝐡}\text{ }p_E.$$ (309) The number of states here is finite. By contrast, in quantum state *estimation* the set of states is infinite. Since a quantum state itself is not observable, quantum state estimation means estimating the density matrix $`\rho `$ of the quantum state, as we have already seen. This, too, can be put in the context of game theory. In the *State Estimation Game* LJ Alice chooses an arbitrary pure state $`|\psi 𝐇_d`$ and sends $`|\psi ^N`$ to Bob and $`|\psi `$ to a referee. After receiving the $`N`$ states from Alice, Bob performs a measurement on them and then sends a pure state $`|\varphi `$ to the referee. After receiving the two states from Bob and Alice, the referee compares them according to some criterion (see cloning, below), then awards a payoff to Alice if the two states are not sufficiently close, or to Bob if they are. Of course Bob’s task is to construct the best quantum state measurement he can given the $`N`$ states received from Alice. ## .20 Quantum cloning In econometrics one tries, by some procedure, to produce an estimate $`\widehat{a}`$ of some unknown parameter $`a`$. This can be considered an attempt, by our estimation procedure, to *clone* the parameter $`a`$. We don’t expect to achieve a perfect clone, but only a best estimate that lies within an interval of uncertainty. Which brings us to the cloning of quantum states. The object of an *optimal cloning device* RFW is to prepare near copies as close to the original as possible. Optimal cloning can be formulated in terms of a quantum game, the *Cloning Game*, played between Alice and Clare, the cloning queen. This game will have $`N`$ input systems and $`M`$ output systems. We start with Alice, who has a pure state described by a density matrix $`\rho `$ in 2-dimensional Hilbert space $`𝐇_2`$. She is going to run her state preparing procedure $`N`$ times, giving rise to a composite system in Hilbert space $`𝐇_2^N`$: $$\mathrm{𝟏}_2^N\rho =\rho ^N.$$ (310) Alice then ships $`\rho ^N`$ off to Clare. Clare uses a cloning device $`T_m`$ of her choice to produce $`M`$ output systems $`T_m\rho ^N`$. Next, Alice produces $`M`$ copies of her original system, $`\rho ^M`$. The outcome of the game depends on $$T_m\rho ^N\text{ vs. }\rho ^M.$$ (311) Since $`T_m`$ maps density matrices to density matrices, it is restricted to being a linear completely positive trace preserving map. One way of assigning payoffs to this game would be to base them on the norm difference $$||T_m\rho ^N\text{ - }\rho ^M||.$$ (312) Another way would be to use the *fidelity*, based on $`\mathrm{𝐭𝐫𝐚𝐜𝐞}(\rho ^MT_m\rho ^N)`$. This would be $`1`$ if the cloning machine were perfect. The fidelity could depend on the input density matrix $`\rho `$. Define $`F(T)`$ by $$F(T)=\text{inf}_\rho \text{ trace }(\rho ^MT_m\rho ^N)<1.$$ (313) Then Clare’s job is to maximize $`F(T)`$. This makes the Cloning Game a maximin problem. A cloner is called ‘universal’ if the fidelity of the output clones is independent of the input state. The maximal fidelity of cloning for a universal cloner is $`\frac{5}{6}`$, which can be achieved by unitary evolution or by a teleportation scheme BDEFMS . A *universal quantum cloner* of 1 qubit $``$ 2 qubits is a quantum machine that takes as input an unknown quantum state $`|\psi `$ and generates as output two qubits in a state that may be described by a density matrix of the form $`\rho =\eta |\psi \psi |+(1\eta )\frac{1}{2}\mathrm{𝟏}`$. The parameter $`\eta `$ describes the shrinking of the original Bloch vector $`𝐫`$ corresponding to the density operator $`|\psi \psi |`$. For example, if $`|\psi \psi |=\frac{1}{2}(\mathrm{𝟏}+𝐫\sigma )`$, then $`\rho =\frac{1}{2}(\mathrm{𝟏}+\eta 𝐫\sigma )`$. Then the optimal cloner involves maximizing the fidelity by maximizing $`\eta <1`$: $$\text{max}_\eta \text{ }F=\psi |\rho |\psi =\frac{1}{2}(1+\eta ).$$ (314) A Bloch vector shrinkage of $`\eta =\frac{2}{3}`$ corresponds to the maximal fidelity of $`\frac{5}{6}`$. The cloning process goes like this. Let $`|B`$ denote the initial state of blank copies (the destination of the clones) plus any auxillary qubits (‘ancilla’) needed in the process. The qubit $`|\psi `$ to be cloned is encoded in the basis $`(|0,|1)`$. Then the universal quantum cloning machine (UQCM) transformation $`T_{UQCM}`$ performs the following transformations on the basis vectors or states: $`T_{UQCM}|0|B\sqrt{{\displaystyle \frac{2}{3}}}|0|0|A_{}+\sqrt{{\displaystyle \frac{1}{6}}}(|01+|10)|A`$ (315) $`T_{UQCM}|1|B\sqrt{{\displaystyle \frac{2}{3}}}|1|1|A+\sqrt{{\displaystyle \frac{1}{6}}}(|01+|10)|A_{}.`$ (316) Here $`A`$ and $`A_{}`$ represent two possible orthogonal final states for the ancilla qubits. Note that this implies for the input state $`|\psi `$, the output $$T_{UQCM}|\psi |B$$ (317) $$(\sqrt{\frac{2}{3}}|0|0|A_{}+\sqrt{\frac{1}{6}}(|01+|10)|A,\text{ }\sqrt{\frac{2}{3}}|1|1|A+\sqrt{\frac{1}{6}}(|01+|10)|A_{})\left(\begin{array}{c}a\\ b\end{array}\right).$$ (318) The next step is to *trace* over the ancilla qubits, which yields a two-qubit mixed state. Then another trace is performed with respect to each individual qubit, giving two copies of the same mixed one-qubit state, which has a fidelity of $`\frac{5}{6}`$ when compared to the original state. ## .21 Conclusion At this point the reader has enough background to start doing quantum game thory. Of course, there is much more to be said, as the references will indicate. The reader is referred especially to the notes on quantum computation EHI ZM JPres . This essay has demonstrated that traditional game theory is a subset of quantum game theory, and the latter has a much richer structure and a broader set of outcomes. That is all the justification required for doing *quantum* game theory. Nothing is given up, and more is obtained by switching to the latter. Therefore the study of traditional game theory is neither an evolutionarily stable strategy nor a Nash equilibrium, and will be relegated to the dust-bin of extinct species and nonequilibrium payoffs. That being said, can the current state of quantum game theory survive an invasion of mutants? I hope those invading mutants will be mathematical economists coming to fix what’s wrong with quantum mechanics. Indeed, Lambertini LL argues that mathematical economics and quantum mechanics are isomorphic. A quantum game $`\mathrm{\Gamma }=\mathrm{\Gamma }(𝐇,\mathrm{\Lambda },U,\{s_i\}_j,\{\pi _i\}_j)`$, where $`𝐇`$ is a Hilbert space; $`\mathrm{\Lambda }`$ is the initial state of the game; $`U`$ is a unitary matrix applied to all the player’s qubits at the beginning and end of the game; $`\{s_i\}_j`$ are the set of moves of player $`j`$, including convex combinations; and $`\{\pi _i\}_j`$ are the set of payoffs to player $`j`$. The purpose of the game is to endogenously determine the strategies that maximize player $`j`$’s expected payoff. Generally, a pure quantum move $`s_i`$ is a unitary matrix applied to the player’s individual qubit. In the course of this essay, we have seen the Spin Flip game, the Guess a Number games I and II, the RSA game, Prisoner’s dilemma, Battle of the sexes, Newcomb’s game, Evolutionarily stable strategy game, Coin flip game, Pseudo-telepathy game, and game theoretic aspects of Teleportation, Secret sharing, State estimation, and Quantum cloning. In the Spin Flip game, Bob was able to exploit quantum superposition via the Hadamard transform $`H`$ to always win the game, though to be sure this outcome was also dependent on the sequence of player moves. The key to Guess a Number Game I was use of the Grover search algorithm to rotate a state vector in Hilbert space to the approximate location of the unknown number. This search was speeded up from $`N`$ moves to $`\sqrt{N}`$ moves by the use of superposition and calls to the $`f_a`$ oracle. In the Guess a Number game II, the Bernstein-Vazirani oracle was used to create the Walsh transform $`W_{2^n}`$ of the unknown number after a single call to the oracle. In the RSA game, Shor’s factoring algorithm was used to project a superimposed state of integers into, with high probability, a number that is near an integer multiple of $`\frac{2^{2n}}{r}`$ for the given composite RSA prime $`N=pq`$, where $`r`$ is the order of the tested element. The probability was controlled by use of the quantum Fourier transform. In the Prisoner’s dilemma game, we saw that the addition of quantum moves $`H`$ and $`\sigma _z`$ to $`\mathrm{𝟏}`$ and $`\sigma _x`$ added to the traditional game outcomes, and indeed attained a Pareto optimal point as a Nash equilibrium. In the Battle of the sexes game, the same quantum moves produced a unique Nash and Pareto optimal equilibrium in pure strategies; and equality between Alice and Bob, also a Nash equilibrium and Pareto optimal, in mixed strategies. Newcomb’s paradox was resolved by the Superior Being’s ability to perfectly predict (control) Alice’s choice through the use of superposition, which replaced omniscience on the part of the Superior Being, and the incentive to cheat on the part of Alice. These games also show, through the use of the unitary matrix U, the partial irrelevance of the categories ‘cooperative’ and ‘noncooperative’. If players’ qubits are entangled in the game, there are hidden channels of communication (an invisible hand) when a player simply focuses on maximizing his or her own expected utility. In the Evolutionarily stable strategy game, invading mutants playing quantum moves were able to wipe out existing species playing only classical moves. The Coin flip game demonstrated the use of a quantum oracle, in a game without entanglement, to turn an unfair game into a fair one. In the Pseudo-telepathy game, communication among players was not necessary in order for them to conspire to win the game, as long as they shared a quantum entangled state. The game could be won with certainty with an implied coalition of all $`N`$ players, while any proper subset of $`N`$ had expected payoff of $`0`$. We also saw that *N-dimensional probability space* was the *core* of the pseudo-telepathy game. Does this mean quantum entanglement gives rise to quantum probability? We saw that qubit states are unobservable, and under measurement are projected onto the measurement basis, typically $`0`$ or $`1`$, and hence destroyed. This creates opportunity as well as difficulties. Measurement in the Bell basis is at the heart of the teleportation protocol. And while quantum states can only be cloned with a certain fidelity, they can be used for secret sharing and secure communication. The problems of quantum state discrimination using maximum likelihood in a Bayesian framework, or quantum state estimation using the same in connection with the Bloch sphere representation of the density matrix, are not concepts fundamentally foreign to economists. Piotrowski and Sladkowski PS3 have stated what they called the *Quantum anthropic principle*: Even if at earlier stages of civilization markets were governed by classical laws, the incomparable efficiency of quantum algorithms in conveying comparative advantage should result in market evolution such that quantum behaviors will prevail over classical ones. Since nature already plays quantum games, it would appear that humans do so also using their personal quantum computers (human brains). Thus, while speculative, Gottfried Mayer’s comment in *Complexity Digest* is not so far fetched: ‘It might be that while observing the due ceremonial of everyday market transactions we are in fact observing capital flows resulting from quantum games eluding classical description. If human decisions can be traced to microsopic quantum events one would expect that nature would have taken advantage of quantum computation in evolving complex brains. In that sense one could indeed say that quantum computers are playing their market games according to quantum rules.’ GJM
warning/0506/astro-ph0506412.html
ar5iv
text
# Propagation of UHECRs ## 1 Introduction In the past ninety years of cosmic rays research there has been a constant search for the *end of the cosmic-rays spectrum* and it has long been thought that this end would be determined by the highest energy that cosmic accelerators might be able to achieve. Despite the continuous search, no end of the spectrum was found. In 1966, right after the discovery of the cosmic microwave background (CMB), it was understood $`^\mathrm{?}`$ that high energy protons would interact inelastically with the photons of the CMB and produce pions. For homogeneously distributed sources this would cause a flux suppression, called the *GZK cutoff*: for the first time the end of the cosmic ray spectrum was related to a physical process rather than to speculations on the nature of the accelerators. Moreover, for the first time, the end of the cosmic ray spectrum was predicted to be at a rather well defined energy, around $`10^{20}\mathrm{eV}`$, where the so-called photo-pion production starts to be kinematically allowed. UHECRs can be of various nature and during their propagation over cosmological distances they suffer different kinds of energy losses. In this paper we consider most of the particles that could play the role of UHECRs and we review the processes affecting their propagation. In §2 we discuss the propagation of protons, heavy nuclei, photons and neutrinos. In §3 we show that the two largest experiments operating up to now, AGASA and HiRes, have a too small statistic to provide a conclusive answer about the presence or absence of the GZK feature in the UHECRs spectrum. We conclude in §4. ## 2 Propagation of UHECRs in the cosmic photon background ### 2.1 Protons There are three sources of energy loss for ultra high energy protons propagating over cosmological distances: the expanding universe redshift, pair production ($`p\gamma pe^+e^{}`$) and photo-pion production ($`p\gamma \pi N`$), each successively dominating as the proton energy increases. For protons the most important background is the CMB and the most important process is the photo-pion production in which a nucleon of sufficiently high energy sees, in its reference frame, the photons of the CMB blue-shifted to $`\gamma `$-rays above the threshold energy for photo-pion production, $`E_\gamma ^{\mathrm{lab},\mathrm{thr}}=m_\pi +m_\pi ^2/(2m_N)160\mathrm{MeV}`$. The cross-section for this process has a pronounced resonance just above threshold, corresponding to the production of an intermediate state $`\mathrm{\Delta }^+`$ that immediately decays into a nucleon and a pion, whereas in the limit of high energies it increases logarithmically with $`s=m_N^2+2m_NE_\gamma ^{\mathrm{lab}}`$, giving rise to multiple pion production. For a background photon of energy $`ϵ`$ in the cosmic rest frame, defined as the frame in which the CMB is isotropic, the threshold energy $`E_\gamma ^{\mathrm{lab},\mathrm{thr}}`$ translates into a corresponding threshold for the nucleon energy: $$E_{\mathrm{thr}}=\frac{m_\pi }{(1\mathrm{cos}\theta )ϵ}\left(m_N+\frac{m_\pi }{2}\right)6.810^{19}\left(\frac{10^3\mathrm{eV}}{ϵ}\right)\left(\frac{2}{1\mathrm{cos}\theta }\right)\mathrm{eV}.$$ (1) Typical CMB photon energies are of the order of $`10^3\mathrm{eV}`$, giving a threshold value of a few tens of EeV (for a head-on collision). The interplay of this threshold with the Planck spectrum of the CMB photons produces a very steep, exponential, curve for the interaction length. The combination with the large inelasticity of the photo-pion interaction (the mean inelasticity goes from $`0.13`$ at threshold to $`0.5`$ at high energy, with large fluctuations) creates a very efficient and rapid mechanism to reduce the nucleon energy and makes the universe opaque to nucleons with energy above $`10^{20}\mathrm{eV}`$ on scales above $`100\mathrm{Mpc}`$. The so-called GZK cutoff is due exactly to this: the flux at earth of nucleons with energy below threshold, say $`510^{19}\mathrm{eV}`$, is due to contributions from (almost) all the universe, from Fig. 2 the loss length at this energy is of the order of $`1\mathrm{Gpc}`$, while doubling the energy the loss length is reduced to $`100\mathrm{Mpc}`$, and only a small portion of the universe contributes to the flux. Thus this change by a factor two in the energy changes the loss length by almost an order of magnitude, which translates in about the same ratio between the flux below $`510^{19}\mathrm{eV}`$ and above $`10^{20}\mathrm{eV}`$ if the sources have no luminosity evolution and no local overdensity and there is no magnetic field. Below $`610^{19}\mathrm{eV}`$ the dominant loss mechanism for protons becomes the production of electron-positrons pairs on the CMB, $`p\gamma pe^+e^{}`$, down to the corresponding threshold: $$E_{\mathrm{thr}}=\frac{m_e}{ϵ}\left(m_N+m_e\right)4.810^{17}\left(\frac{10^3\mathrm{eV}}{ϵ}\right)\mathrm{eV}.$$ (2) The interaction length for this process is much shorter than the one for pion production, but on the other hand the inelasticity is much lower, $`10^3`$. This makes the pair production loss length of the order of Gpc (see Fig. 2). The low inelasticity of pair production allows in calculations to treat this process as a continuous energy loss, whereas the pion production has to be treated as a discrete process due to its large inelasticity. The last important mechanism which dominates near and below the pair production threshold is redshifting due to the expansion of the universe. Fig. 2 shows the loss lengths for pion and pair production as calculated in Ref. $`^\mathrm{?}`$. It is worth stressing that what has been named the GZK cutoff is in fact a *feature* $`^\mathrm{?}`$ as the shape of the energy spectrum around $`10^{20}\mathrm{eV}`$ depends on many unknowns. The modifications of the spectrum shape due to the above-mentioned loss processes was first investigated by Berezinsky and Grigorieva in Ref. $`^\mathrm{?}`$. They calculated the modification factor (basically the observed spectrum divided by the injection spectrum) for a uniform distribution of sources up to a maximum distance $`d_{\mathrm{max}}`$. Fig. 2 shows their results for sources without cosmological evolution, $`m=0`$, for some values of the maximum distance of the sources. For large $`d_{\mathrm{max}}`$, which is the case we are interested in, the spectrum shows a steepening followed by a flattening and then by a suppression. The flattening is due to the interplay between the features produced by the pair and pion production processes and it is an important feature for these spectra since it has a characteristic shape. There are claims that this feature has been observed in the experimental data $`^\mathrm{?}`$, although it is not yet clear if the feature in the data is due to this effect or if it is due to the transition between the galactic and extra-galactic components. It is important to stress what we said above: what is generically called GZK-cutoff is actually a *feature* as the spectrum does not end at $`10^{20}\mathrm{eV}`$ (see Fig. 2), but has a flux suppression that depends on many details such as the injection spectrum of cosmic rays, the luminosity evolution of the sources, the local overdensity of sources and the magnetic field strength in the intergalactic medium. As an example, including the luminosity evolution makes the sources at high redshift brighter that the nearby ones and this enhances the flux suppression, while a local overdensity of sources has the opposite effect $`^\mathrm{?}`$; a flatter spectrum produces a lesser attenuation than a steeper one and the strength of the magnetic field in the intergalactic medium con produce many interesting features, see for example Ref. $`^\mathrm{?}`$. ### 2.2 Heavy Nuclei For nuclei the situation is slightly different: the dominant loss process above about $`10^{19}\mathrm{eV}`$ is photodisintegration in the CMB and IR background (IRB) due to the giant dipole resonance, followed at lower energy by the pair production. The photo-pion production process is negligible, except for light nuclei at very high energies $`^{\mathrm{?},\mathrm{?}}`$. Indeed, for a nucleus of mass number $`A`$ and charge $`Z`$, the energy loss length for pion production is roughly the same one of a nucleon with identical Lorentz factor. This is due to the fact that the cross section for pion production is approximately proportional to the mass number $`A`$, while the inelasticity is proportional to $`1/A`$. For pair production we got a different behavior because, while the inelasticity is proportional to $`1/A`$ as before, the cross section is proportional to $`Z^2`$ resulting in an energy loss length lower by a factor $`A/Z^2`$ with respect to a proton with the same Lorentz factor. Since $`ZA/2`$, the ratio of the photo-pair and photo-pion production increases roughly linearly with $`Z`$ $`^\mathrm{?}`$. The cross sections for photodisintegration $`\sigma _{A,i}(ϵ^{})`$ contains essentially two regimes depending on $`ϵ^{}`$, the photon energy in the nucleus rest frame. At $`ϵ^{}<30\mathrm{MeV}`$ there is the domain of the giant dipole resonance and the disintegration proceeds mainly by the emission of one or two nucleons. At higher energies, the cross section is dominated by multi-nucleon emission for heavy nuclei and is approximately flat up to $`ϵ^{}150\mathrm{MeV}`$. A useful quantity to estimate the energy loss rate by photodisintegration is given by the effective rate: $$R_A^{\mathrm{eff}}=\frac{\mathrm{d}A}{\mathrm{d}t}=\underset{i}{}iR_{A,i}.$$ (3) For photodisintegration, the average fractional energy loss results equal to the fractional loss in mass number of the nucleus, $`E^1\mathrm{d}E/\mathrm{d}t=A^1\mathrm{d}A/\mathrm{d}t`$, because the nucleon emission is isotropic in the rest frame of the nucleus. Therefore during the photodisintegration process the Lorentz factor of the nucleus is conserved, unlike the cases of pair and pion production which involve the creation of new particles that carry away energy. The energy loss time for photodisintegration is then $`A/R_A^{\mathrm{eff}}`$. Fig. 3 shows separately the different contributions to this quantity from CMB, IR and optical photons for Fe nuclei, together with the total one (solid line) and the pair creation loss length. It is apparent that the optical background has no relevant effect, that the IR one dominates the photodisintegration processes below $`10^{20}\mathrm{eV}`$ and the CMB dominates above $`10^{20}\mathrm{eV}`$. The pair creation rate is relevant for Fe energies $`410^{19}\mathrm{eV}÷210^{20}\mathrm{eV}`$ ($`\gamma `$ factors $`(1÷4)10^9`$), for which the typical CMB photon energy in the rest frame of the nucleus is above threshold ($`>1\mathrm{MeV}`$) but still well below the peak of the giant resonance ($`10÷20\mathrm{MeV}`$). The effect of pair creation losses is to reduce the $`\gamma `$ factor of the nucleus, obviously leaving $`A`$ unchanged $`^\mathrm{?}`$. We should not get fooled by the loss lengths in Fig. 3 into thinking that if the loss length for a Fe nucleus of $`10^{20}\mathrm{eV}`$ is $`500\mathrm{Mpc}`$, then we can receive on Earth a Fe nucleus that started many hundred Mpc away. This is because the corresponding interaction length is more than an order of magnitude shorter and after every interaction the nucleus becomes lighter and lighter and along with this the loss length for photodisintegration becomes shorter and shorter. The net result $`^\mathrm{?}`$, as can be seen in Fig. 5, is that after $`10\mathrm{Mpc}`$ all the energies are below $`210^{20}\mathrm{eV}`$ and after $`100\mathrm{Mpc}`$ they are below $`10^{20}\mathrm{eV}`$. ### 2.3 Photons As in the case of UHE nucleons and nuclei, the propagation of UHE photons (and electrons/positrons) is also governed by their interaction with the cosmic photon background. The dominant interaction processes in this case are the attenuation of UHE photons due to pair production (PP) on the background photons ($`\gamma \gamma _be^+e^{}`$), and inverse Compton scattering (ICS) of the electrons (positrons) on the background photons. The $`\gamma `$-ray threshold energy for PP on a background photon of energy $`ϵ`$ is $$E_{\mathrm{thr}}=\frac{m_e^2}{ϵ}2.610^{11}\left(\frac{ϵ}{\mathrm{eV}}\right)^1\mathrm{eV},$$ (4) whereas ICS has no threshold. In the high-energy limit, the total cross sections for PP and ICS are: $$\sigma _{\mathrm{PP}}2\sigma _{\mathrm{ICS}}\frac{3}{2}\sigma _\mathrm{T}(m_e^2/s)\mathrm{ln}(s/2m_e^2)(sm_e^2).$$ (5) For $`sm_e^2`$, $`\sigma _{\mathrm{ICS}}`$ approaches the Thomson cross section $`\sigma _\mathrm{T}=8\pi \alpha ^2/3m_e^2`$ ($`\alpha `$ is the fine structure constant), whereas $`\sigma _{\mathrm{PP}}`$ peaks near the threshold. Therefore, the most efficient targets for electrons and $`\gamma `$-rays of energy $`E`$ are background photons of energy $`ϵm_e^2/E`$. For UHE this corresponds to $`ϵ<10^6\mathrm{eV}100\mathrm{MHz}`$. Thus, radio background photons play an important role in UHE $`\gamma `$-ray propagation through extragalactic space $`^\mathrm{?}`$. Unfortunately, the universal radio background (URB) is not very well known, mostly because it is difficult to disentangle the Galactic and extragalactic components. In the extreme Klein-Nishina limit, $`sm_e^2`$, either the electron or the positron produced in the process $`\gamma \gamma _be^+e^{}`$ carries most of the energy of the initial UHE photon. This leading electron can then undergo ICS whose inelasticity (relative to the electron) is close to 1 in the Klein-Nishina limit. As a consequence, the upscattered photon which is now the leading particle after this two-step cycle still carries most of the energy of the original $`\gamma `$-ray, and can initiate a fresh cycle of PP and ICS interactions. This leads to the development of an *electromagnetic (EM) cascade* which plays an important role in the resulting observable $`\gamma `$-ray spectra. An important consequence of the EM cascade development is that the effective penetration depth of the EM cascade, which can be characterized by the energy attenuation length of the leading particle (photon or electron/positron), is considerably greater than just the interaction length (see Fig. 5$`^\mathrm{?}`$. EM cascades play an important role particularly in some exotic models of UHECR origin such as collapse or annihilation of topological defects in which the UHECR injection spectrum is predicted to be dominated by $`\gamma `$-rays. But, even if only UHE nucleons and nuclei are produced in the first place, for example via conventional shock acceleration, EM cascades can be produced by the secondaries coming from the decay of pions which are created in interactions of UHE nucleons with the low energy photon background $`^\mathrm{?}`$. Most of the energy of fully developed EM cascades ends up below $`100\mathrm{GeV}`$ where it is constrained by measurements of the diffuse $`\gamma `$-ray flux. Flux predictions involving EM cascades are therefore an important source of constraints of UHE energy injection on cosmological scales. It should be mentioned that the development of EM cascades depends sensitively on the strength of the extragalactic magnetic fields (EGMFs) which is rather uncertain. The EGMF typically inhibits the cascade development because of the synchrotron cooling of the $`e^+e^{}`$ pairs produced in the PP process. The energy lost through synchrotron radiation does not, however, disappear; rather, it reappears at lower energies and can even initiate fresh EM cascades. ### 2.4 Neutrinos The propagation of UHE neutrinos is governed mainly by their interactions with the relic neutrino background (RNB). The interaction energies are typically smaller than electroweak energies even for UHE neutrinos and then the cross sections are given by the Standard Model of electroweak interactions which are well confirmed experimentally. Physics beyond the Standard Model is not expected to play a significant role in UHE neutrino interactions with the low-energy relic backgrounds. Despite the neutrino-neutrino cross section are at least a few order of magnitude smaller than the neutrino-nucleon ones, the latter interactions are negligible compared to interactions with the RNB because the RNB particle density, $`100\mathrm{cm}^3`$ per family, is about 10 orders of magnitude larger than the baryon density. The $`\nu \overline{\nu }`$ annihilation mean free path is of the order of $`\lambda _\nu =(n_\nu \sigma _{\nu \overline{\nu }})^1410^{28}\mathrm{cm}`$, just above the present size of the horizon ($`H_0^110^{28}\mathrm{cm}`$). The neutrino is the only known stable particle that can propagate through the universe essentially uninhibited even at the highest energies. This has lead to the speculation that neutrinos could be indeed the super-GZK primaries. However, in the Standard Model a neutrino incident vertically in the atmosphere would pass through it uninhibited, never initiating an extensive air shower. Consequently, for these scenarios to work, one has to postulate new interactions so that these neutrinos acquire a strong cross section above $`10^{20}\mathrm{eV}`$. An interesting situation arises if the RNB consists of massive neutrinos with $`m_\nu 1\mathrm{eV}`$: such neutrinos would constitute hot dark matter which is expected to cluster, for example, in galaxy clusters. This would potentially increase the interaction probability for any neutrino of energy within the width of the $`Z^0`$ resonance at $`E=M_Z^2/2m_\nu =410^{21}(\mathrm{eV}/m_\nu )\mathrm{eV}`$. It has been suggested that the stable end products of the *Z-bursts* induced at close-by distances ($`<50\mathrm{Mpc}`$) from Earth may explain the highest energy cosmic rays $`^\mathrm{?}`$. The problem with these proposals is however that they require a very high flux of UHE neutrinos to begin with and this makes Z-burst above GZK energies more likely to play a role in the context of non-accelerating scenarios. For further information see Ref. $`^\mathrm{?}`$ and references therein. It is important to point out that the only conventional/assured source of UHE neutrinos is the GZK effect itself. The neutrinos are the result of the decay of the pions produced in the $`p\gamma `$ interaction. The flux however is not very high and the detection is quite difficult. For further informations see Refs. $`^{\mathrm{?},\mathrm{?}}`$. ## 3 AGASA and HiRes: is there a discrepancy? AGASA and HiRes are, up to now, the two experiments with the larger exposure for the detection of UHECRs. They reported however apparently conflicting results. The two reported spectra appear: 1) to have a systematic offset at low energy and 2) to differ above $`10^{20}\mathrm{eV}`$ where AGASA shown no hint of the GZK-suppression whereas HiRes seems to be consistent with it. It has been shown $`^\mathrm{?}`$ that a systematic overestimate of the AGASA energies by 15% and a corresponding underestimate of the HiRes energies by the same amount would in fact bring the two data sets in a much better agreement in the region below $`10^{20}\mathrm{eV}`$. In Ref. $`^\mathrm{?}`$ we applied our Monte Carlo simulation $`^{\mathrm{?},\mathrm{?},\mathrm{?}}`$ to investigate the discrepancies at high energy and we found that: * assuming a uniform distribution of sources, the AGASA spectrum is reproduced, in a conventional scenario where the average spectrum has the GZK suppression, with a probability of $`610^4`$ ($`3\sigma `$). * assuming the presence of the 15% systematic error, the shifted AGASA data are reproduced with a probability of $`610^3÷10^2`$ ($`2.5\sigma ÷2\sigma `$). * the HiRes data are reproduced, in a scenario without a GZK suppression<sup>a</sup><sup>a</sup>aTo mimic a scenario without GZK suppression we used the AGASA dataset as template., with a probability of $`2\%`$, ($`2\sigma `$). It is important to stress that in order to properly do these calculations one has to take into account the statistical error in the energy determination. Due to the steeply falling spectrum and to the expected change of slope around $`10^{20}\mathrm{eV}`$ the statistical error in the energy determination changes the average expected number of events above $`10^{20}\mathrm{eV}`$ by $`1`$ and with the present limited statistics even a difference of one event is important $`^\mathrm{?}`$. In Fig. 6 we plot the spectra of some of the simulated AGASA realizations that produced 11 or more events above $`10^{20}\mathrm{eV}`$. It is striking the resemblance of the simulated spectra to the AGASA one: all of them show no evidence of the GZK suppression. This shows that the AGASA spectrum is far from being impossible, even if the average cosmic ray spectrum can be expected to show a GZK feature. From the above points we can conclude that neither AGASA nor HiRes have enough statistical power to prove the presence or absence of the GZK feature in the spectrum of UHECRs. A new generation of experiments is needed to finally provide a conclusive answer to this question. ## 4 Conclusions We considered several particles that can possibly play the role of UHECRs and we discussed the energy loss processes that affect their propagation over cosmological distances. We showed that for all considered particles the energy losses above $`10^{20}\mathrm{eV}`$ become so severe that they cannot propagate over distances larger than about $`100\mathrm{Mpc}`$ without reducing their energy below $`10^{20}\mathrm{eV}`$. The effect of these energy losses on the spectrum of UHECRs is the so called GZK-suppression, due to the fact that below $`10^{20}\mathrm{eV}`$ almost all the universe is contributing to the observed flux whereas above $`10^{20}\mathrm{eV}`$ we receive contributions only from sources not too far away, $`100\mathrm{Mpc}`$. We stressed the point that this is not a cutoff, but a feature since the spectrum does not end at $`10^{20}\mathrm{eV}`$, but it is only suppressed, and the amount of this suppression depends on many unknowns such as: the luminosity evolution of the sources, their local overdensity and the magnetic field strength in the intergalactic medium. We showed that the present sets of data are not enough to determine whether this GZK suppression is present or not in the observed spectrum and that we need a new generation of experiments to have a conclusive answer to this question. ## Acknowledgments This research is funded in part by NASA grant NAG5-10919. ## References
warning/0506/nucl-ex0506013.html
ar5iv
text
# Photoproduction at Hadron Colliders ## 1 Introduction The upcoming Large Hadron Collider (LHC) will reach proton-proton energies an order of magnitude higher than any existing accelerator. Because relativistic protons and heavier nuclei accompanied by fields of virtual photons, the LHC can be used to study photonuclear and two-photon interactions at energies far beyond those accessible at HERA or other accelerators. Photoproduction is of interest in both pp and heavy-ion collisions. Proton-proton collisions produce photons with the highest energies, and, because of the very high $`pp`$ luminosities, good rates. However, for many channels, the signal to noise ratio may be lower than in ion collisions. Heavy ions are accompanied by very high photon fluxes, and, because of the very strong fields, a single ion-ion collision can induce multiple electromagnetic interactions. The correlations between the multiple photons lead to the ability to “tune” the photon beam, by selecting different photon energy spectra and polarizations. Since these reactions take place at large impact parameters, where no hadronic interactions occur, they are often known as “ultra-peripheral collisions” (UPCs). UPCs can be used to study a variety of topics ourreview baurreview kraussreview . Low$`x`$ gluon distributions can be probed via heavy quark (including quarkonium) and jet production. UPCs can be used for many other studies of QCD. At the LHC, photonuclear interactions can be used to search for new physics. The strong fields allow for many tests of quantum electrodynamics in the very strong field regime, where perturbation theory may be expected to fail. Many of these topics are already being studied at RHIC. ## 2 Photoproduction at Hadron Colliders For most reactions, the photon flux from protons or nuclei is well described by the Weizsäcker-Williams method of virtual photons. The photon flux per unit area for an energy $`\omega `$ at a distance $`b`$ from a relativistic nucleus with charge $`Z`$ is jackson $$N(\omega ,b)=\frac{Z^2\alpha \omega ^2}{\pi ^2\gamma ^2\mathrm{}^2}K_1^2(x)$$ (1) where $`x=\omega b/\gamma `$, $`\gamma `$ is the Lorentz boost of the nucleus, $`\alpha 1/137`$ is the fine structure constant, and $`K_1`$ is a modified Bessel function. The total photon flux from an ion with radius $`R_A`$ is $$n(\omega )=d^2bN(\omega ,b).$$ (2) The constrant $`b>R_A`$ is usually imposed to eliminate the photon flux inside the nucleus (where Eq. 1 fails, and, in any case, most of the flux is not usable). For photonuclear or two-photon interactions to be visible, the two nuclei must not interact hadronically, requiring $`b>2R_A`$. This flux is calculated numerically, but can be approximated within about 15% by requiring $`b>2R_A`$ in Eq. 2 ourreview . The cross section for photonuclear interactions can be written BKN $$\sigma (A+AA+A+X)=d^2bP(b)$$ (3) where $`P(b)`$ is the probability for a photonuclear interaction, $`P(b)=𝑑\omega N(\omega ,b)\sigma _{\gamma A}(\omega )`$ where $`\sigma _{\gamma A}(\omega )`$ is the cross section for the photonuclear interaction in question. This formulation is easy to generalize to include multiple interactions between a single ion pair: $$\sigma (A+AX_1+X_2+\mathrm{})=d^2P_1(b)P_2(b).$$ (4) In general, $`P(b)1/b^2`$, so the integrand for a $`n`$photon reaction goes as $`1/b^{2n}`$ and the more photons involved in a reaction, the smaller the average impact parameters factorize . For example, with gold at RHIC, the median impact parameter drops from 46 fm for unselected $`\rho ^0`$ production to 18 fm for $`\rho ^0`$ accompanied by mutual Coulomb excitation BKN . The smaller impact parameters harden the photon spectrum, from $`1/\omega `$ to independent of $`\omega `$. For some reactions, $`P(2R_A)>1`$; in this case $`P(b)`$ is the mean number of reaction at that $`b`$. Factorization can be used to simplify triggering on UPCs. One reaction can serve as a ’trigger’ for another. STAR has studied the reactions $`Au+AuAu^{}+Au^{}+\rho ^0`$ and $`Au+AuAu^{}+Au^{}+e^+e^{}`$, using signals from the neutrons emitted in the $`Au^{}`$ decays to trigger the detector, providing $`\rho ^0`$ and $`e^+e^{}`$ samples without trigger bias. The photon polarizations are also correlated. The electric field of the photon-emitting nucleus parallels the impact parameter vector, so photons are linearly polarized along the impact parameter vector. For multiple interactions between a single ion-pair, the parallel polarizations can lead to observable angular correlations between decay products factorize . ## 3 Results from STAR at RHIC The STAR collaboration has produced final results on $`\rho ^0`$ production STARrho and on two-photon production of $`e^+e^{}`$ pairs STARee . Events were selected with two types of triggers: minimum bias triggers that select events with mutual Coulomb dissociation, taking advantage of factorization, and topological triggers, that select low multiplicity events with appropriate topologies in the central detector STARrho . The $`\rho `$ data is well described by the soft Pomeron model, and the previously discussed factorization holds. In the soft Pomeron model, the incident photon fluctuates to a quark-antiquark pair, which then elastically scatters (via Pomeron exchange) from the target nucleus KN99 . Because the scattering is coherent, the momentum transfer is limited to order $`\mathrm{}/R_A`$. This low $`p_T`$ is a distinctive experimental signature; for gold, most of the signal occurs for $`p_T<150`$ MeV/c. Figure 1 shows the $`t_{}=p_T^2`$ spectrum of $`\rho ^0`$ with rapidity $`0.1<|\eta |<0.6`$, selected with stringent cuts to minimize the background STARinterfere . At moderate and high $`t`$, the spectrum is well fit by an exponential, $`dN/dt=a\mathrm{exp}(bt)`$. However, for $`t<0.0015`$GeV<sup>2</sup>, the data drops off. This drop can be explained by interference between two indisinguishable possibilities: nucleus 1 emits a photon which interacts with nucleus 2, or vice-versa theoryinterfere . In $`pp`$ or $`AA`$ collisions, these two possibilities are related by a parity transformation. Since the $`\rho ^0`$ is negative parity, the interference is destructive. At mid-rapidity, $$\sigma =\sigma _0\left[1\mathrm{cos}(p_Tb)\right]$$ (5) Of course, $`b`$ is unknown, and the overall interference depends on the integral over all $`b`$. Away from $`y=0`$, the interference is reduced because the photon energies, fluxes, amplitudes etc. for the two directions are different. The solid curve in Fig. 1 shows a fit to a functional form based on these factors; for this sample, the interference is $`101\pm 8(\mathrm{stat}.)\pm 15(\mathrm{syst}.)\%`$ of that expected STARinterfere . Because the two sources are spatially separated, the final state $`\pi ^+\pi ^{}`$ wave function does not factorize into single-particle wave functions, and the system exhibits the Einstein-Podolsky-Rosen paradox EPR . For $`\overline{p}p`$ collisions, the transformation between the two possibilies is a charge-parity transformation; vector mesons are $`CP`$ positive, so the interference in Eq. 5 is positive ppinterfere ; this may be studied at the Fermilab Tevatron. STAR has also studied $`\rho ^0`$ production in $`dAu`$ collisions. The photon is usually emitted by the gold nucleus, and the deuteron is the target. Both coherent (deuteron stays intact) and incohent (deuteron dissociates) interactions have been observed. The $`t_{}`$ spectrum for the incoherent interactions is similar to that observed in $`eA`$ collisions at HERA timser . The STAR $`e^+e^{}`$ data is well described by lowest order quantum electrodynamics and factorization STARee . The $`p_T`$ spectrum of the $`e^+e^{}`$ pairs is not well described by the virtual photon paradigm - the photon virtuality is required to fit the data. STAR has also studied 4-prong final states, like $`\pi ^+\pi ^{}\pi ^+\pi ^{}`$. Fig. 2 compares the $`p_T`$ spectrum of 4-prong events with net charge 0 with those of net charge 2. This data was taken in 2002 with the minimum-bias trigger. A neutral excess is present for $`p_T<150`$ MeV/c, with the mass spectrum of the excess centered around 1.5 GeV/c<sup>2</sup>. This work was supported by the U.S. D.O.E. under Contract No. DE-AC-03076SF00098.
warning/0506/cond-mat0506531.html
ar5iv
text
# Intrinsic Spin Hall Edges ## Abstract The prediction of intrinsic spin Hall currents by Murakami et al. and Sinova et al. raised many questions about methods of detection and the effect of disorder. We focus on a contact between a Rashba type spin orbit coupled region with a normal two-dimensional electron gas and show that the spin Hall currents, though vanishing in the bulk of the sample, can be recovered from the edges. We also show that the current induced spin accumulation in the spin orbit coupled system diffuses into the normal region and contributes to the spin current in the leads. Transport and manipulation of spins in semiconductor structures has become a mainstream in condensed matter physics Zutic . In principle, spins can be injected into semiconductors by ferromagnets via electric contacts. However, finding suitable material combinations that do not suffer from the conductance mismatch Schmidt , turned out to be difficult. Furthermore, introducing ferromagnetic materials into the semiconductor microfabrication process is undesirable from a technological point of view. The prospect to generate spin accumulations in semiconductors without ferromagnets or applied magnetic fields simply by driving a current through a material with intrinsic spin-orbit (SO) interaction and broken inversion-symmetry Levitov ; Edelstein ; Katoaccum is therefore very attractive. A related effect that attracted a lot of attention is the spin Hall effect (SHE), i.e. the spin current(SC) that has been predicted to flow normal to an applied electric current in the absence of an applied magnetic field. When caused by impurities with spin-orbit scattering DP ; extrinsic this effect is called “extrinsic”. A spin Hall current (SHC) can also be generated by the spin-orbit interaction of the lattice potential as has recently been predicted for p-doped III-V semiconductors Murakami and the two-dimensional electron gas with a Rashba-type SO interaction (R2DEG) Sinova . Whether the experimental observations of the spin Hall effect by optical methods exp have intrinsic or extrinsic origin is still a matter of debate. In spite of initial controversies, analytic theories Inoue ; pdoped ; Mishchenko ; Khaetskii ; Chalaev as well as numerical simulations Nomura ; Sheng consistently predict that the SHE should vanish in the disordered (bulk) R2DEG SHE\_Workshop . Some doubts remain whether the SC, being a non-conserved quantity in SO coupled systems, is observable at all Rashba . In this Letter, we focus on the spin currents near normal contacts. First, an elementary and general proof is given that the spin Hall effect due to the lattice SO coupling (viz. intrinsic SHE) must vanish in diffuse bulk systems with an arbitrarily strong SO interaction that is linear in the electron wave vector. Nevertheless, using an extension of this argument to finite system sizes, we show that near the edges a spin Hall current can persist. Next, by solving the kinetic equations for a model system of a R2DEG in contact with a normal metal system without SO interaction, we calculate indeed a finite SHC. This SC is generated in a skin depth determined by the Dyakonov-Perel DP spin-flip diffusion length ($`L_s`$) and the polarization is not normal to the 2DEG, having a component due to the diffusion current from the SO-generated spin accumulation (SA). The magnitude of the SC generated at the edges depends on whether the system is clean (impurity broadening less than the SO splitting) or dirty (opposite limit). However, in contrast to the bulk SC, the edge SC does not vanish when the system is not ballistic ($`L_s`$ smaller than the system size). The SC is calculated in the normal metal contact and therefore certainly a transport current Rashba . Related work on interface and boundary effects focused so far on mesoscopic systems via numerical simulations Nikolic and the SA near hard wall boundaries hardwall . We proceed to derive a transport equation valid in the Boltzmann limit that is capable of handling the full spin dynamics. In $`2\times 2`$ spin space, the Hamiltonian is $$H=𝐩^2/2m+V(𝐱)+H^Re𝑬(t)𝐱,$$ (1) where $`𝐱`$ and $`𝐩`$ are the (two-dimensional) position and momentum operators, respectively. Here the unit vector $`\widehat{𝒛}`$ is normal to the 2DEG, $`H^R=(\alpha /\mathrm{})𝐩(𝝈\times \widehat{𝒛})`$ is the Rashba Hamiltonian with Pauli matrices $`𝝈`$ and $`\alpha `$ parameterizes the strength of the SO interaction Nitta , $`𝑬`$ is the electric field, and $`V(𝐱)=_{i=1}^N\varphi (𝐱𝑿_i)`$ is the impurity potential, modelled by $`N`$ impurity centers located at points $`\{𝑿_i\}`$. Although it is possible to consider ac fields, we focus here on dc fields in the $`x`$direction and assume that the electric field is turned on adiabatically in the remote past at which the system was in thermal equilibrium, i.e. we assume $`𝑬=lim_{s0}𝑬_0\mathrm{exp}(st/\mathrm{})`$. To leading order in the impurity potential the diagonal elements of the density matrix in reciprocal space satisfy the following equation Luttinger : $`isf(𝒌)+[f(𝒌),H_k^R]`$ $`={\displaystyle \underset{k^{}}{}}\left(f_{𝒌𝒌^{}}V_{𝒌^{}𝒌}V_{𝒌𝒌^{}}f_{𝒌^{}𝒌}\right)`$ $`+e𝑬[𝒙,f^0].`$ (2) Here $`f^0=(f_+^0+f_{}^0)/2+\sigma _\theta (f_+^0f_{}^0)/2`$ is the equilibrium density matrix with $`f_\pm ^0(k)=\left(\mathrm{}^2k^2/2m\pm \alpha k\right)`$, where $`(E)`$ is the Fermi function and $`\sigma _\theta =𝐤(𝝈\times 𝒛)/k`$. The off-diagonal elements of the density matrix read $$f_{𝒌𝒌^{}}=\frac{i}{2\pi }_{\mathrm{}}^{\mathrm{}}𝑑EG_𝒌^R(E)\left(f(𝒌)f(𝒌^{})\right)G_𝒌^{}^A(E)V_{𝒌𝒌^{}}.$$ (3) Here $`G_k^{R(A)}(E)=(EH_0H_k^R+()is/2)^1`$ are retarded (advanced) matrix Green functions. Substituting Eq. (3) into Eq. (2) and averaging over $`V_{kk^{}}`$ (in the Boltzmann limit averaging is equivalent to replacing $`|V_{kk^{}}|^2`$ with its average value $`N|\varphi _{kk^{}}|^2/A^2`$, where $`N`$ is the number of impurities and $`A`$ is the area Luttinger ) gives our basic equation, valid for weak $`V`$ and low enough impurity densities to ignore weak localization effects, but to all orders in $`\alpha `$. The mechanism behind the intrinsic SHE is the spin precession of quasiparticles while being accelerated by the electric field Sinova . However, impurity scattering provides a brake that in the steady state cancels the acceleration on average. Therefore the SHE should vanish in an infinite, homogeneously disordered system. This idea can be formally expressed by considering the acceleration operator $`\ddot{𝐱}_i=e𝑬_i/m_iV/m2\alpha ^2\mathrm{}^3ϵ_{3ji}\sigma _3𝐩_j`$, where $`ϵ_{ijk}`$ is the antisymmetric tensor and the Einstein summation convention is implied. We notice that the last term is proportional to the $`j`$’th component of the SC operator polarized in the $`z`$direction, $`J_j^z=\{𝒗_j,𝝈_z\}`$. The expectation value is defined by $`O\overline{\mathrm{Tr}fO}`$, where the trace is over wave vector and spin space, and $`\overline{\mathrm{}}`$ denotes averaging with respect to impurity configurations. In a steady state the average acceleration $`\ddot{𝐱}_i`$ must vanish, leading to the equality $$2\alpha ^2m^2\mathrm{}^3ϵ_{3ji}J_j^z=e𝑬_i_iV.$$ (4) We show that the right hand side of this equality also vanishes by evaluating the expectation value of the deceleration due to impurity scattering: $`_iV`$ $`=i{\displaystyle \underset{kk^{}}{}}(𝒌_i𝒌_i^{})V_{kk^{}}\mathrm{tr}f_{𝒌^{}𝒌}`$ $`=i{\displaystyle \underset{k}{}}\mathrm{tr}\left(𝒌_i[f(𝒌),H_𝒌^R]+i𝒌_ie𝑬_j_{k_j}f^0\right)=e𝑬_i,`$ where $`\mathrm{tr}`$ is the trace over spin components and Eqs. (2-3) have been used in the second step. Substituting the expression above into Eq. (4) we see that all components of the SC polarized in the $`z`$direction vanish with the average acceleration caveat . This result holds for infinite systems regardless of the range of the impurity potential or whether the system is clean ($`\alpha k_F\tau /\mathrm{}1`$) or dirty ($`\alpha k_F\tau /\mathrm{}1`$), where $`\tau `$ is the momentum lifetime. Thus generalizing previous results Inoue ; pdoped ; Mishchenko ; Khaetskii ; Chalaev . However, as we shall show below, for semi-infinite and finite systems, SCs persist near the edges, but the size of these currents depend on whether the system is clean or dirty. This line of argument allows one to check related Hamiltonians. In the presence of $`k`$-linear Dresselhaus and Rashba terms, the result remains unchanged besides the substitution $`\alpha ^2\alpha ^2\beta ^2`$, where $`\beta `$ is the Dresselhaus spin orbit coupling constant. Thus the SHC still vanishes (with the possible exception of the degeneracy point $`\alpha =\beta `$ FT:dresselhaus ). When the SO coupling contains cubic terms like $`\alpha (k)=\alpha _0+\alpha _1k^2`$, it is easy to show that the SHC is proportional to $`\alpha _1`$ pdoped . Another possible situation is the presence of a Zeeman field: in this case the operator equation is modified to give $`\alpha ϵ_{3ji}J_j^z=\sigma _iB_3\sigma _3B_i`$, relating the SHC to the SA. If $`\alpha `$ varies in space, the SHC is found to be proportional to the spatial derivatives of $`\alpha `$ and $`f`$. If $`\alpha `$ is constant, but $`f`$ varies, e.g. due to boundaries or interfaces Mishchenko , the SHC is proportional to the gradients of the density matrix: $$\frac{4\alpha ^2m^2}{\mathrm{}^3}ϵ_{3ji}\widehat{J}_j^z=k_i\{\frac{\mathrm{}^2k_l}{m}+\alpha (\widehat{𝒛}\times 𝝈)_l,_lf(𝒌,𝒙)\}.$$ (5) This equation shows that although the bulk SH current vanishes, there is no a priori reason for SH currents near the edges of the R2DEG to vanish. Next, we shall show indeed the SH currents do not vanish near the edges. We therefore return to the quantum transport equation, allowing for spatially varying density matrices but assuming short-range s-wave scatterers with $`\overline{|V_{kk^{}}|^2}=N\lambda ^2/A`$ and $`\alpha /k_F1`$. We solve the transport equation by expressing $`f`$ in terms of a gradient expansion of $`\rho (E)=(i\mathrm{}^2/2\pi m)_k\left(G_k^R(E)f(𝒌)f(𝒌)G_k^A(E)\right)`$. In the case of s-wave scatterers and to leading order in $`m\alpha /\mathrm{}^2k_F`$, this generates the same diffusion equation as Ref. Mishchenko and Burkov et al. in Ref. Sinova . In terms of components of the density matrix: $`\rho =n+𝐬\mathbf{}𝝈+\sigma _3s_3`$ : $`D^2n4K_{sc}(\mathbf{}\times 𝐬)_z`$ $`=0`$ (6) $`D^2s_32K_p(\mathbf{}𝐬)`$ $`={\displaystyle \frac{2s_3}{\tau _s}}`$ (7) $`D^2𝐬+2K_p\mathbf{}s_3K_{sc}(𝐳\times \mathbf{})n`$ $`={\displaystyle \frac{𝐬}{\tau _s}}`$ (8) Here $`D=v_F^2\tau /2`$, $`\tau _s=\tau (1+4\xi ^2)/2\xi ^2`$, $`K_{sc}=\alpha \xi ^2/(1+4\xi ^2)`$, $`K_p=\mathrm{}k_F\xi /m(1+4\xi ^2)^2`$ and $`\xi =\alpha k_F\tau /\mathrm{}`$. The SC is given, in the diffuse limit by $$j_j^i=\frac{v_F\xi }{1+4\xi ^2}\left(\delta _{i3}\left(s_jϵ_{jm3}\frac{\alpha \tau }{2}_mn\right)\delta _{ij}s_3\right)D_js_i.$$ (9) Electric field dependence can be reintroduced by the substitution $`\mathbf{}\mathbf{}+e𝐄_E`$. We now focus on a four terminal structure as depicted in Fig. 1. This structure consists of two massive reservoirs biased to produce a charge current in the $`x`$direction. Between the reservoirs there is a R2DEG hybrid structure, with $`\alpha (𝐱)=\alpha _0`$ for $`L>y>0`$ and $`\alpha (𝐱)=0`$ for $`y<0`$ and $`y>L`$, and additional differences are disregarded. The normal 2DEGs are coupled to massive reservoirs R3 and R4, that are biased such that the charge current is zero, but a SC can still be collected. To the order (in $`\alpha /k_F`$) that we are considering, $`n`$ does not depend on $`y`$. Alternatively, one can assume that the transverse size of the leads to R3 and R4 are much smaller than the distance between R1 and R2, in which case one can also neglect the $`y`$ dependence of $`n`$. We consider the distributions at a safe distance from the reservoirs R1 and R2. In this region $`s_y`$ and $`s_z`$ depend only on $`y`$ and $`s_x=0`$ and the diffusion equation becomes: $$\left(\begin{array}{cc}\frac{d^2}{d\overline{y}^2}1& 2\eta \frac{d}{d\overline{y}}\\ 2\eta \frac{d}{d\overline{y}}& \frac{d^2}{d\overline{y}^2}2\end{array}\right)\left(\begin{array}{c}s_2\\ s_3\end{array}\right)=\left(\begin{array}{c}\tau _sK_{sc}\frac{dn}{dx}\\ 0\end{array}\right),$$ (10) where $`\eta =(1+4\xi ^2)^{3/2}`$, $`\overline{y}=y/L_s`$ and $`L_s=\sqrt{D\tau _s}`$. In order to derive the matching condition for the spin and charge distribution functions at the contacts (i.e. interface between the R2DEG and 2DEG), short-range fluctuations of boundaries and interfaces that can lead to additional spin relaxation REF:Vasko are disregarded. We consider an arbitrary solution $`\chi `$ of the Schrödinger equation set by the Hamiltonian Eq. (1). We label the solutions in the R2DEG and 2DEG regions $`\chi _R`$ and $`\chi _N`$, respectively. At the interface $`\chi _R|_0=\chi _N|_0`$ and $`𝐧(i+\alpha _0(𝐳\times \sigma ))\chi _R|_0=i𝐧\chi _N|_0,`$ where $`𝐧`$ is the unit vector normal to the interface. Multiplying from the left with $`\chi |_0^{}\sigma _i`$ and evaluating the imaginary part we obtain: $`𝐧\left(i\chi _N^{}\sigma _i\chi _Ni(\chi _N)^{}\sigma _i\chi _N\right)|_0`$ $`=𝐧(i\chi _R^{}\sigma _i\chi _Ri(\chi _R)^{}\sigma _i\chi _R`$ $`+\alpha _0\chi _R^{}\{\sigma _i,(𝐳\times \sigma )\}\chi _R\left)\right|_0.`$ We identify the right (left) hand side of this equation as the SC density in the Rashba (normal) 2DEG. In terms of spin density matrices we have $`\mathrm{tr}\left(f^R\{\sigma _i,𝐧𝐣(0)\}\right)=\mathrm{tr}\left(f^N\{\sigma _i,𝐧𝐣(0)\}\right)`$, where $`𝐣(𝐱)\{𝐯,\delta (\widehat{𝐱}𝐱)\}`$ is the local current density operator. We therefore have to match the normal components of the SC density given by Eq. (9) at the interface softBC . Since the operator $`\{𝐯_i,𝝈_j\}`$ can have a nonzero expectation value in the equilibrium state it has been questioned whether it governs transport of spins in the presence of SO interaction Rashba . We notice that the negative energy solutions (relative to the band crossing) of the Hamiltonian Eq. (1) without the electric field term, are localized to the R2DEG region if surrounded by a region with $`\alpha =0`$. In a normal 2DEG surrounding the R2DEG, we can therefore show that equilibrium SCs exposed in Ref. Rashba do not transmit into the normal region. Moreover, the expectation value of the SC density operator vanishes for these localized solutions and it is precisely the absence of contributions from these solutions that shifts the equilibrium value of the SC to zero. Returning to the setup in Fig. 1, we assume that the reservoirs R3 and R4 are sufficiently large such that all components of the SA at their respective interfaces with the ordinary 2DEG leads vanish. Shrinking the widths of the 2DEGs to zero, we obtain the effective boundary conditions $`s_i=0`$ at the R2DEG$`|`$R3 interface. The finite Ohmic resistance of a finite 2DEG region between the R2DEG and the reservoir can easily be reintroduced if necessary and would lead to somewhat smaller spin conductances. We then can solve the diffusion equation above and obtain the spin current using Eq. (9). Analytical formulae turn out to be too lengthy to reproduce here. Our results are therefore summarized in Fig. 2. The SA is suppressed at the interface, reflecting the massive-reservoir boundary condition. The gradient of the two components $`s_y`$ and $`s_z`$ leads to two SC components. The SC polarized in the $`y`$-direction represents the out-diffusion of the bulk $`s_y`$ SA. This is not a Hall current, since it flows into the side contacts with opposite directions (Fig. 2) with polarization that is inverted with the bias current direction. The resulting spin conductivity at the interface is $`\sigma _{xy}^y=0.87e\xi ^2/2\pi `$ in the dirty limit($`\xi 1`$). For larger values of $`\xi `$, $`\sigma _{xy}^y`$ increases above this quadratic behavior, but in the clean limit ($`\xi 1`$) this increase is cut off by the resistance of the normal region. When spins diffuse from a finite distance into the 2DEG, they precess in the SO-generated magnetic field. Consequently there is a diffusion current polarized along the $`z`$-direction, for which we find in the dirty limit a conductivity $`\sigma _{xy}^z=0.83e\xi ^2/2\pi `$. The conductivity $`\sigma _{xy}^z`$, contrary to $`\sigma _{xy}^y`$, decreases below this quadratic behavior for larger values of $`\xi `$ and vanishes in the clean limit. Nikolic et alNikolic recently observed SCs with $`z`$ and $`y`$ polarization in numerical simulations. In addition, we also find a SHC exponentially localized to the edges that decays in the bulk on the length scale $`L_s`$ and reaches its maximum value $`eEK_{sc}/2\pi \alpha `$ at the interface to the reservoir. This is due to the fact that the first term in Eq. (9) being proportional to $`s_y`$(thus zero at the interface) is no longer screening the second term (proportional to $`n`$) and reflects the physical process that the SHC density generated near the interface can escape into the reservoir before it decays due to spin relaxation. The resulting dc spin Hall conductivity is given by $`e\xi ^2/2\pi `$ in the dirty and $`e/8\pi `$ in the clean limit. Our result differs from that of Ref. Mishchenko who did not take into account the edge currents and obtained similar values only for the ac response at carefully tuned frequencies. In conclusion, we find that in a Hall geometry two different spin currents can be extracted by the Hall contacts from the current-biased disordered R2DEG. In addition to the SHC, the current-induced SA drives a spin-diffusion current. The SO generated spin accumulation is therefore not confined to the region where it is generated, but can be extracted and, at least in principle, used as a source of spins for spintronics applications. Both diffusion and SHCs are generated within a strip that scales like the Dyakonov-Perel spin diffusion length. We thank Junichiro Inoue, Philip Stamp, Fei Zhou and especially Yuli V. Nazarov for useful discussions. This work was supported by the FOM, EU Commission FP6 NMP-3 project 05587-1 “SFINX”, NSERC Canada discovery grant number R8000 and PITP.
warning/0506/math0506612.html
ar5iv
text
# Automorphisms of K3 surfaces ## 1. Introduction We work over the complex numbers field $``$. A smooth projective surface $`X`$ is a $`K3`$ surface if the canonical line bundle (is trivial) $`\omega _X𝒪_X`$ and if the irregularity $`h^1(X,𝒪_X)=0`$. In this note we shall report some progress we made recently (mostly with K. Oguiso) on automorphisms of $`K3`$ surfaces. By abuse of notation, we denote by the same $`\omega _X`$ a nowhere vanishing global holomorphic $`2`$-form on $`X`$. So $`H^0(X,\omega _X)=\omega _X`$. Now the Hodge decomposition says that $`H^2(X,)=\omega _XH^{1,1}(X)\overline{\omega _X}`$. Let $`\text{Aut}(X)`$ be the group of all automorphisms of a $`K3`$ surface $`X`$. Take a subgroup $`G\text{Aut}(X)`$. For every $`g`$ in $`G`$, we have $`g^{}\omega _X=\alpha (g)\omega _X`$ for some $`\alpha (g)`$ in the multiplicative group $`^{}=\{0\}`$. One sees that $`\alpha :G^{}`$ is a group homomorphism. Let $`G_N=\text{Ker}(\alpha )`$ be the kernel. By Nikulin , Theorem 0.1 or Sterk , Lemma 2.1, the image $`\text{Im}(\alpha )`$ is a finite subgroup (of order $`I=I(G)=I(X,G)`$) in $`^{}`$, so it equals the multiplicative group $`\mu _I=\zeta _I`$, where $`\zeta _I=\mathrm{exp}(2\pi \sqrt{1}/I)`$ is an $`I`$-th primitive root of $`1`$. Then one has the basic exact sequence of groups: $$1G_NG\mu _I1.$$ One may say that the study of $`G`$ is reduced to that of $`G_N`$ and $`\mu _I`$, where $`G_N`$ acts on $`X`$ symplectically (see Notation and Terminology below) while $`h^{}\omega _X=\zeta _I\omega _X`$ for some pre-image $`h`$ in $`G`$ of a generator in $`\mu _I`$. ###### Problem 1.1. \[Extension Problem\] Given symplectic subgroup $`H`$ and a cyclic subgroup $`\mu _I`$ of $`\text{Aut}(X)`$ on a $`K3`$ surface $`X`$, find the conditions on $`H`$ and $`I`$ such that there is a subgroup $`G`$ of $`\text{Aut}(X)`$ with $`G_N=H`$ and $`I=I(X,G)`$. In the forthcoming paper Ivanov-Oguiso-Zhang , we will tackle this problem. Now a summary of the contents. We start with symplectic actions on $`K3`$ surfaces, then purely non-symplectic actions, and actions of mixed type. Finally we mention the connection between $`K3`$ surfaces with a non-symplectic action and log Enriques surfaces initially defined in Zhang and . Examples of actions are either included or mentioned about their whereabouts. In §3, we also provide a new proof of the impossibility of $`/(60)`$ acting purely non-symplectically on a $`K3`$ surface, by using Lefschetz fixed point formula for vector bundles. Notation and Terminology For a $`K3`$ surface $`X`$, we use the same symbol $`\omega _X`$ ($`𝒪_X`$) to denote the canonical line bundle and also a generator of $`H^0(X,\omega _X)=\omega _X`$. Let $`S_X=\text{Pic}(X)<H^2(X,)`$ be the Neron Severi lattice and $`T_X=(S_X)^{}=\{tH^2(X,)|(t,s)=0,sS_X\}`$ the transcendental lattice. A group $`G`$ is called a $`K3`$ group if $`G`$ is a subgroup of $`\text{Aut}(X)`$ for some $`K3`$ surface $`X`$. An element $`g`$ of $`\text{Aut}(X)`$ is simplectic if $`g^{}\omega _X=\omega _X`$. A group $`G`$ is symplectic if every element of $`G`$ is symplectic. An element $`g`$ (or the group $`g`$) of order $`n`$ is purely non-symplectic if $`g^{}\omega _X=\eta _n\omega _X`$ for some primitive $`n`$-th root of $`1`$. For an integer $`I2`$, the Euler function $`\phi (I)=\mathrm{\#}\{n|\mathrm{\hspace{0.17em}1}nI1;\mathrm{gcd}(n,I)=1\}`$. Set $`\zeta _I=\mathrm{exp}(2\pi \sqrt{1}/I)`$, a primitive $`I`$-th root of $`1`$. Set $`\mu _I=\zeta _I`$, a cyclic group of order $`I`$. For groups $`H`$ and $`K`$, the symbol $`H.K`$ denotes a group $`G`$ such that $`H`$ is normal in $`G`$ and $`G/HK`$, i.e., $`G`$ is an extension of $`K`$ by $`H`$. When $`K`$ can be chosen to be a subgroup of $`G`$, then we denote $`H:K`$, i.e., $`G`$ is a split extension of $`K`$ by $`H`$. $`2^n`$, $`3^n`$, etc denote elementary abelian groups $`(/(2))^n`$, $`(/(3))^n`$, etc. $`S_n`$ is the symmetric group (of order $`n!`$) in $`n`$ letters. $`A_n`$ is the alternating group (of order $`n!/2`$) in $`n`$ letters. $`M_{24},M_{23},M_{20},M_{10}`$, etc are Mathieu groups. $`Q_8`$ is the quaternion group of order 8, while $`D_n`$ is the dihedral group of order $`n`$ (see Mukai ). Acknowledgement. The author would like to thank Professors S. Kondo and K. Oguiso for the valuable comments. ## 2. Symplectic $`K3`$ groups The classification of abelian symplectic groups was done by Nikulin . Especially, he proved: ###### Theorem 2.1. Let $`g`$ be a symplectic automorphism on a $`K3`$ surface $`X`$. Then $`\text{ord}(g)8`$. Moreover, the fixed locus (point wise) $`X^g`$ has cardinality $`8,6,4,4,2,3`$, or $`2`$, if $`\text{ord}(g)`$ equals $`2,3,4,5,6,7`$, or $`8`$, respectively. Mukai has determined all maximum finite symplectic $`K3`$ groups (11 of them); see also Kondo for a lattice-theoretic proof: ###### Theorem 2.2. For a finite group $`G`$, the following are equivalent: $`(1)`$ $`G`$ has a (faithful) symplectic action on a $`K3`$ surface. $`(2)`$ $`G`$ has an embedding $`GM_{23}`$ into the Mathieu group and decomposes $`\{1,2,3,\mathrm{},24\}`$ into at least $`5`$ orbits. (The bigger Mathieu group $`M_{24}`$ acts naturally on the $`24`$-element set with the smaller $`M_{23}`$ as a one-point stabilizer subgroup.) $`(3)`$ $`G`$ has a Mathieu representation $`V`$ over Q with $`dimV^G5`$ and the $`2`$-Sylow subgroups of $`G`$ can be embedded into $`M_{23}`$. $`(4)`$ $`G`$ is isomorphic to one of the $`11`$ (maximum symplectic) groups below where the order is indicated in the parenthesis: $`\text{PSL}_2(7)`$ $`(168)`$, $`A_6`$ $`(360)`$, $`S_5`$ $`(120)`$, $`M_{20}=2^4:A_5`$ $`(960)`$, $`F_{384}=2^4:S_4`$ $`(384)`$, $`A_{4,4}=(S_4\times S_4)A_8`$ $`(288)`$, $`T_{192}=(Q_8Q_8):S_3`$ $`(192)`$, $`H_{192}=2^4:D_{12}`$ $`(192)`$, $`N_{72}=3^2:D_8`$ $`(72)`$, $`M_9=3^2:Q_8`$ $`(72)`$, $`T_{48}=Q_8:S_3`$ $`(48)`$. The result above completely classifies all finite symplectic actions on $`K3`$ surfaces. The possible subgroups $`G`$ (exactly 80 of them) of the 11 maximum symplectic $`K3`$ groups acting on some $`K3`$ surface $`X`$ together with the data of $`\text{Sing}(X/G)`$ has been given by Xiao , the list. ## 3. Purely non-symplectic $`K3`$ groups We now turn to the purely non-symplectic $`K3`$ groups. Suppose that a cyclic group $`\mu _I=g`$ ($`I2`$) acts on a $`K3`$ surface $`X`$ such that $`g^{}\omega _X=\eta _I\omega _X`$, where $`\eta _I`$ is a primitive $`I`$-th root of 1. By Nikulin Theorem 0.1, or Sterk Lemma 2.1, the Euler function $`\phi (I)|\text{rank}(T_X)`$. Since $`\text{rank}(T_X)=22\text{rank}(S_X)21`$, we get: ###### Lemma 3.1. Suppose $`g`$ is an order-$`I`$ ($`I2`$) automorphism of a $`K3`$ surface $`X`$ such that $`g^{}\omega _X=\eta _I\omega _X`$ with $`\eta _I`$ a primitive $`I`$-th root of $`1`$. Then $`\phi (I)21`$ and hence $`I`$ equals one of the following: Case $`\phi (I)=20`$. $`I=66,50,44,33,25`$. Case $`\phi (I)=18`$. $`I=54,38,27,19`$. Case $`\phi (I)=16`$. $`I=60,48,40,34,32,17`$. Case $`\phi (I)=12`$. $`I=42,36,28,26,21,13`$. Case $`\phi (I)=10`$. $`I=22,11`$. Case $`\phi (I)=8`$. $`I=30,24,20,16,15`$. Case $`\phi (I)=6`$. $`I=18,14,9,7`$. Case $`\phi (I)=4`$. $`I=12,10,8,5`$. Case $`\phi (I)=2`$. $`I=6,4,3`$. Case $`\phi (I)=1`$. $`I=2`$. Conversely, each $`I`$ with $`I60`$ in the lemma above is geometrically realizable (see Kondo , or Machida-Oguiso , Proposition 4\]). ###### Proposition 3.2. For each $`I`$ with $`I60`$ in the lemma above, there is a $`K3`$ surface $`X`$ such that $`\text{Aut}(X)\mu _I=g`$ and $`g^{}\omega _X=\zeta _I\omega _X`$, where $`\zeta _I=\mathrm{exp}(2\pi \sqrt{1}/I)`$ is a primitive $`I`$-th root of 1. Indeed, $`\mu _{60}`$ can not act purely non-symplectically by the result below (the early proofs are Machida-Oguiso , Theorem 5.1\] and Xiao , Theorem, page 1); our proof here comes from the very simple idea: the calculation of Lefschetz number by Atiyah, Segal and Singer, , , for the trivial line bundle, and hence more straightforward. Therefore, the lemma and proposition above give the complete classification for the $`\mu _I`$ appearing in the basic exact sequence in the Introduction. ###### Proposition 3.3. (see and ). There is no $`K3`$ surface with a purely non-symplectic action by the group $`\mu _{60}`$. Namely, there is no pair $`(X,g)`$ of a $`K3`$ surface $`X`$ and an order $`60`$ automorphism $`g`$ with $`g^{}\omega _X=\zeta _{60}\omega _X`$. ###### Proof. Suppose the contrary that there is a pair $`(X,g)`$ as in the statement. Set $`\omega =\omega _X`$ and $`\zeta =\zeta _{60}`$, so $`g^{}\omega =\zeta \omega `$. Let $`\{C_i\}`$ be the set of all $`g`$-fixed (point wise) irreducible curves. So at a point $`Q`$ on $`C_i`$, one can diagonalize $`g^{}`$ as $`(1,\zeta )`$ with appropriate coordinates. Let $`n=(1g(C_i))`$; we set $`n=0`$ when there is no such $`g`$-fixed curve $`C_i`$. Let $`M_j=\{P_{jk}\}`$ be the set of points $`P`$ at which $`g^{}`$ can be diagonalized as $`(\zeta ^j,\zeta ^{j+1})`$ with $`1j<60/2=30`$. Set $`m_j=|M_j|`$ and $`m=m_j`$. Then the fixed locus $`X^g`$ (point wise) is the disjoint union of smooth curves $`C_i`$ and $`m`$ isolated points $`P_{jk}`$ (see, e.g., Oguiso - Zhang , Lemma 2.1 (3) for the smoothness). By Atiyah - Segal and Atiyah - Singer at pages 542 and 567, we can calculate the Lefschetz number $`L(g)`$ in two different ways: $$L(g)=\underset{i=0}{\overset{2}{}}(1)^i\text{Tr}(g^{}|H^i(X,𝒪_X)),$$ $$L(g)=\underset{j,k}{}a(P_{jk})+\underset{i}{}b(C_i),$$ $$a(P)=\frac{1}{det(1g^{}|T_P)},$$ $$b(C_i)=\frac{1g(C_i)}{1\zeta }\frac{\zeta C_i^2}{(1\zeta )^2}=\frac{(1g(C_i))(1+\zeta )}{(1\zeta )^2}.$$ Here $`T_P=T_{X,P}`$ is the tangent space at $`P`$, and $`\zeta ^1`$ is the eigenvalue of the action $`g_{}`$ on the normal bundle of $`C_i`$. We see then that $`_ib(C_i)=n(1+\zeta )/(1\zeta )^2`$. Also $`a(P_{jk})=_jm_j/(1\zeta ^j)(1\zeta ^{j+1})`$. On the other hand, by the Serre duality $`H^2(X,𝒪_X)=H^0(X,\omega )^{}`$, we have $`L(g)=1+\zeta ^1`$. Identify the two $`L(g)`$, so $`x:=\zeta `$ satisfies the equation below: $$F_1(x)=(1+x^1)+\underset{j=1}{\overset{29}{}}\frac{m_j}{(1x^j)(1x^{j+1})}+\frac{n(1+x)}{(1x)^2}=0.$$ Note that the minimal polynomial $`\mathrm{\Phi }_{60}(x)`$ of $`x=\zeta `$ over Q is given as follows; this is also obtained by factorizing $`x^{30}+1`$: $$\mathrm{\Phi }_{60}(x)=x^{16}+x^{14}x^{10}x^8x^6+x^2+1.$$ Our $`x=\zeta `$ also satisfies the following equations (factorizing $`(x^{10})^3+1`$): $$x^{30}+1=0=x^{20}x^{10}+1.$$ Since $`x^{\pm 30}=1`$, we have $`(1x^i)(1x^{i+1})=(1+x^{30i})(1+x^{i+130})`$. Substituting these equalities for $`15i29`$ into the equation $`F_1(x)=0`$, we obtain a new equation $`0=F_1(x)=F_2(x)/G_2(x)`$, where $`F_2(x)`$ is a polynomial of degree $`146`$ in $`x`$ with coefficients in $`m_j`$, and $`G_2(x)`$ is the product of polynomials each of which is of degree $`12`$ in $`x`$. Since the minimal polynomial $`\mathrm{\Phi }_{60}(x)`$ of $`x=\zeta `$ is of degree $`16`$, our $`x=\zeta `$ satisfies $`G_2(x)0`$ and $`F_2(x)=0`$. Substituting $`x^{30k}=1`$ (resp. $`1`$) when $`k`$ is even (resp. odd) into the equation $`F_2(x)=0`$, we obtain a polynomial equation $`F_3(x)=0`$ of degree $`29`$ in $`x`$. Substituting $`x^{20}=x^{10}1`$ into the equation $`F_3(x)=0`$, we obtain an equation $`F_4(x)=0`$ of degree $`19`$ in $`x`$. Finally substituting $`\mathrm{\Phi }_{60}(x)=0`$ into the equation $`F_4(x)=0`$, we obtain an equation $`F_5(x)=_{i=0}^{15}d_ix^i=0`$ of degree $`15`$ in $`x`$, where each $`d_i`$ is an integer linear combination of $`m_j`$’s. Note that $`x^i=\zeta ^i`$ ($`0i15`$) are linearly independent over Q. So we get linear equations in $`m_j`$: $`d_0=d_1=\mathrm{}=d_{15}=0`$. Solving them, we arrive at the following relations: $$()4m_1=1+2m_2+2m_{20}2m_{21}+4m_{22}4m_{23}+2m_{26}$$ $$2m_{27}+4m_{28}4m_{29}2m_3+4m_64m_7+2m_82m_9+8n,$$ $$9m_{10}=6012m_{17}+18m_{18}17m_{19}141m_294m_{20}+81m_{21}189m_{22}$$ $$+198m_{23}+6m_{24}12m_{25}81m_{26}+81m_{27}156m_{28}+156m_{29}$$ $$+63m_36m_424m_5198m_6+189m_7105m_8+96m_9+312n,$$ $$36m_{11}=26748m_{17}+108m_{18}104m_{19}582m_2394m_{20}+342m_{21}792m_{22}$$ $$+828m_{23}+24m_{24}48m_{25}342m_{26}+342m_{27}624m_{28}+624m_{29}$$ $$+270m_324m_496m_5828m_6+792m_7438m_8+402m_9+1248n,$$ $$36m_{12}=57+12m_{17}72m_{18}+56m_{19}282m_2134m_{20}+162m_{21}252m_{22}$$ $$+216m_{23}24m_{24}+12m_{25}162m_{26}+162m_{27}384m_{28}+384m_{29}$$ $$+90m_384m_4+24m_5288m_6+252m_7138m_8+102m_9+768n,$$ $$72m_{13}=39+48m_{17}144m_{18}+104m_{19}426m_2182m_{20}+234m_{21}252m_{22}$$ $$+180m_{23}60m_{24}+84m_{25}270m_{26}+234m_{27}600m_{28}+600m_{29}$$ $$+126m_3156m_4+60m_5324m_6+252m_7138m_8+102m_9+1200n,$$ $$18m_{14}=15+12m_{17}36m_{18}+32m_{19}48m_22m_{20}+$$ $$18m_{21}18m_{23}24m_{24}+30m_{25}36m_{26}+36m_{27}96m_{28}$$ $$+87m_{29}30m_4+24m_5+18m_636m_7+6m_86m_9+174n,$$ $$18m_{15}=30+12m_{17}36m_{18}+32m_{19}66m_256m_{20}+72m_{21}144m_{22}$$ $$+126m_{23}+12m_{24}6m_{25}54m_{26}+54m_{27}96m_{28}+87m_{29}$$ $$+18m_3+6m_412m_5126m_6+108m_748m_8+48m_9+174n,$$ $$72m_{16}=39+96m_{17}8m_{19}+138m_2+86m_{20}90m_{21}+180m_{22}$$ $$180m_{23}12m_{24}+60m_{25}+54m_{26}90m_{27}+168m_{28}168m_{29}$$ $$54m_3+12m_4+12m_5+180m_6180m_7+138m_8102m_9336n.$$ To double check that there is no any human error in the above substitutions, we substitute the above relations back into the original equation and get $`F_1(x)=\mathrm{\Phi }_{60}(x)F_6(x)/G_6(x)`$, where $`F_6(x)`$ is a polynomial of degree $`130`$ in $`x`$ and $`G_6(x)`$ is a product of polynomials each of which is of degree $`12`$ in $`x`$. So no error should have occurred during the substitution process, noting that $`x=\zeta `$ satisfies its minimal polynomial (over Q) $`\mathrm{\Phi }_{60}(x)=0`$. The actual simple linear algebra calculation is done by the ”Maple”, though a calculation by hand is possible if the reader is patient and careful enough to avoid any error in doing the simple arithmetics and Gaussian elimination. Now to conclude the proof, we note that the first equality (\*) in the final relations above says that $`4m_1=1+2s`$, with $`s`$ an integer (a integer linear combination of $`m_i`$ and $`n`$). This is impossible because $`m_1`$ is an integer. We reach a contradiction. So there is no such pair $`(X,\mu _{60})`$ and the proposition is proved. ∎ ###### Remark 3.4. As pointed out by Oguiso (and done in ), indeed, one can say a bit stronger than the proposition above (consider $`X/G_N`$): If $`X`$ is a $`K3`$ surface and $`GAut(X)`$ is finite then $`\mu _{I(X,G)}\mu _{60}`$. This is because one always has $`G_N=(1)`$ and $`G=\mu _{I(X,G)}`$ whenever $`\phi (I(X,G))10`$ and $`I(X,G)28`$; see Ivanov-Oguiso-Zhang , or Machida-Oguiso the proof of Lemma (4.1). For a $`K3`$ surface $`X`$, the cohomology $`H^2(X,)=U^3E_8^2`$ is an even lattice, where $`U=e_1+e_2`$ with $`e_i^2=0`$ and $`e_1.e_2=1`$, and $`E_8`$ is the negative definite even lattice of rank 8. So the study of $`\text{Aut}(X)`$ is reduced to that of $`\text{Aut}(S_X)`$ to some extent. Indeed, by Pjateckii-Shapiro and Shafarevich (see also Burns - Rapoport ) up to finite index and finite co-index, $`\text{Aut}(X)`$ is isomrphic to $`\text{Aut}(S_X)/\mathrm{\Gamma }(X)`$ where $`\mathrm{\Gamma }(X)`$ is generated by reflections associated to $`(2)`$-vectors. Let us consider the natural exact sequence $$(1)H_X\text{Aut}(X)\text{Aut}(S_X).$$ Note that $`H_X`$ acts faithfully on $`T_X`$ and on the 1-dimensional space $`\omega _X`$ (see Nikulin Cor. 3.3, or Sterk Lemma 2.1). Indeed, $`H_X`$ is a cyclic finite group of order $`h_X`$ with the Euler function $`\phi (h_X)|\text{rank}(T_X)`$. The results below show that these $`X`$ with maximum $`\phi (h_X)=rk(T_X)`$ are unique up to isomorphisms. For the unique examples of the $`X`$’s in the results below, we refer to Kondo ; see also Examples 3.7 and 3.10 below. ###### Theorem 3.5. (see Kondo , Main Th). Set $`\mathrm{\Sigma }:=\{66,44,42,36,28`$, $`12\}`$. $`(1)`$ Let $`X`$ be a $`K3`$ surface with $`\phi (h_X)=\text{rank}(T_X)`$ and with unimodular transcendental lattice $`T_X`$. Then $`h_X\mathrm{\Sigma }`$. $`(2)`$ Conversely, for each $`N\mathrm{\Sigma }`$, there exists, modulo isomorphisms, a unique $`K3`$ surface $`X`$ such that $`h_X=N`$ and $`\phi (h_X)=\text{rank}(T_X)`$. Moreover, $`T_X`$ is unimodular for this $`X`$. ###### Theorem 3.6. (see Oguiso - Zhang , Theorem $`2`$). Set $`\mathrm{\Omega }:=\{3^k(1k3),\mathrm{\hspace{0.17em}5}^{\mathrm{}}(\mathrm{}=1,2),\mathrm{\hspace{0.17em}7},11,13,17,19\}`$. $`(1)`$ Let $`X`$ be a $`K3`$ surface with $`\phi (h_X)=\text{rank}(T_X)`$ and with non-unimodular transcendental lattice $`T_X`$. Then $`h_X\mathrm{\Omega }`$. $`(2)`$ Conversely, for each $`N\mathrm{\Sigma }`$, there exists, modulo isomorphisms, a unique $`K3`$ surface $`X`$ such that $`h_X=N`$ and $`\phi (h_X)=\text{rank}(T_X)`$. Moreover, $`T_X`$ is non-unimodular for this $`X`$. For $`I=19,17`$, or $`13`$, the unique pair is given in Example 3.7 below. By Theorem 2.1 and Lemma 3.1, every prime order $`p`$ automorphism $`g`$ of a $`K3`$ surface $`X`$ satisfies $`p19`$. Let us give examples for the largest three $`p`$ (see Kondo ). ###### Example 3.7. Here are examples of pairs $`(X_p,\mu _p)`$ ($`p=19,17`$ or 13) of the $`K3`$ surface $`X_p`$ with a $`\mu _p=g_p`$ action, given by its Weierstrass equation and the action on the coordinates. $`X_{19}:y^2=x^3+t^7x+t`$, $`g_{19}^{}(x,y,t)=(\zeta _{19}^7x,\zeta _{19}y,\zeta _{19}^2t)`$. $`X_{17}:y^2=x^3+t^7x+t^2`$, $`g_{17}^{}(x,y,t)=(\zeta _{17}^7x,\zeta _{17}^2y,\zeta _{17}^2t)`$. $`X_{13}:y^2=x^3+t^5x+t^4`$, $`g_{13}^{}(x,y,t)=(\zeta _{13}^5x,\zeta _{13}y,\zeta _{13}^2t)`$. Indeed, these pairs are unique with no conditions on $`h_X`$ or so on; but they turn out to satisfy the conditions in the above theorem: ###### Theorem 3.8. (Oguiso-Zhang Cor. $`3`$). Let $`X`$ be a $`K3`$ surface with an automorphism $`g`$ of order $`p=19`$, $`17`$ or $`13`$ (the three largest possible prime orders. Then the pair $`(X,g)`$ is isomorphic to the pair $`(X_p,g_p)`$ in the above example. ###### Remark 3.9. A similar result as above does not hold for $`p=11`$. Indeed, there are three 1-dimensional families of pairs $`(X_t,\mu _{11})`$ of $`K3`$ surfaces $`X_t`$ with a $`\mu _{11}`$ action; see Oguiso-Zhang . Now we give Kondo’s examples of purely non-symplectic $`\mu _I`$ action on $`K3`$ surfaces with $`I`$ a composite (see Kondo , or Oguiso-Machida , Prop. 4). ###### Example 3.10. Here are examples of pairs $`(X_I,\mu _I)`$ (the first few bigger $`I`$ in Lemma 3.1) of the $`K3`$ surface $`X_I`$ with a $`\mu _I=g_I`$ purely non-symplectic action, given by its Weierstrass equation or as weighted hypersurface and the action on the coordinates; see Kondo . $`X_{66}:y^2=x^3+t(t^{11}1)`$, $`g_{66}^{}(x,y,t)=(\zeta _{66}^{40}x,\zeta _{66}^{27}y,\zeta _{66}^{54}t)`$. $`X_{50}=\{z^2=x_0^6+x_0x_1^5+x_1x_2^5\}(1,1,1,3)`$, $`g_{50}^{}[x_0:x_1:x_2:z]=[x_0:\zeta _{25}^{20}x_1:\zeta _{25}x_2:z]`$. $`X_{44}:y^2=x^3+x+t^{11}`$, $`g_{44}^{}(x,y,t)=(\zeta _{44}^{22}x,\zeta _{44}^{11}y,\zeta _{44}^{34}t)`$. $`X_{54}:y^2=x^3+t(t^91)`$, $`g_{54}^{}(x,y,t)=(\zeta _{27}^2x,\zeta _{27}^3y,\zeta _{27}^6t)`$. $`X_{38}:y^2=x^3+t^7x+t`$, $`g_{38}^{}(x,y,t)=(\zeta _{19}^7x,\zeta _{19}y,\zeta _{19}^2t)`$. $`X_{48}:y^2=x^3+t(t^81)`$, $`g_{48}^{}(x,y,t)=(\zeta _{48}^2x,\zeta _{48}^3y,\zeta _{48}^6t)`$. Indeed, these surfaces (and the action to some extent) are unique: ###### Theorem 3.11. (Machida-Oguiso , Main Th $`1`$). Suppose that $`X`$ is a $`K3`$ surface with a finite group $`G\text{Aut}(X)`$ such that $`I=I(X,G)\{66,50,44`$, $`33,25\}`$ (i.e., $`\phi (I)=20`$). $`(1)`$ The pair $`(X,G)`$ is isomorphic to the pair $`(X_I,g_I)`$ in the example above if $`I`$ is even, and $`(X,G)`$ is isomorphic to the pair $`(X_{2I},g_{2I}^2)`$ if $`I`$ is odd. $`(2)`$ We have $`\text{Aut}(X)=G\mu _I`$ when $`I`$ is even, and $`\text{Aut}(X)\mu _{2I}`$ when $`I`$ is odd. ###### Remark 3.12. A slight weak result holds for $`I=54,38,48`$ (with the unique surface $`X_I`$); see Xiao , Theorem, page 1. So far, we have considered automorphisms of $`K3`$ surfaces with bigger order. We now consider the opposite cases. We start with some examples. ###### Example 3.13. Let $`\zeta =\zeta _4=\sqrt{1}`$. Let $`E=E_\zeta =/(+\zeta )`$ be the elliptic curve of period $`\zeta `$. Let $`X_2\overline{X}_2:=(E\times E)/\text{diag}(\zeta ,\zeta )`$ be the minimal resolution of the quotient surface $`\overline{X}_2`$. Then $`X_2`$ is a $`K3`$ surface. Let $`g_2`$ be the automorphism of $`X_2`$ induced by the action $`\text{diag}(1,1)`$ on $`E\times E`$. Then $`g_2^{}\omega _{X_2}=\omega _{X_2}`$ and the fixed locus (point wise) $`X^{g_2}`$ consists of 10 smooth rational curves (see Oguiso-Zhang , Example 2 for details). ###### Example 3.14. Let $`\zeta =\zeta _3`$. Let $`E=E_\zeta =/(+\zeta )`$ be the elliptic curve of period $`\zeta `$. Let $`X_3\overline{X}_3:=(E\times E)/\text{diag}(\zeta ,\zeta ^2)`$ be the minimal resolution of the quotient surface $`\overline{X}_3`$. Then $`X_3`$ is a $`K3`$ surface. Let $`g_3`$ be the automorphism of $`X_3`$ induced by the action $`\text{diag}(\zeta ,1)`$ on $`E\times E`$. Then $`g_3^{}\omega _{X_3}=\zeta \omega _{X_3}`$ and the fixed locus (point wise) $`X^{g_3}`$ consists of 6 smooth rational curves and 9 isolated points (see Oguiso-Zhang , Example 1 for details). From the result above or intuitively, we see that a $`K3`$ surface $`X`$ with a non-symplectic action of bigger order is uniquely determined by the action. The result below shows that the other extreme case may happen too. ###### Theorem 3.15. (Oguiso-Zhang , Theorem $`4`$). Let $`(X,g)`$ be a pair of a $`K3`$ surface $`X`$ and an automorphism $`g`$ of $`X`$ satisfying (where the conditin $`(3)`$ is removable by Nikulin , or Zhang Th $`3`$): $`(1)`$ $`g^2=id`$, $`(2)`$ $`g^{}\omega _X=\omega _X`$, $`(3)`$ The fixed locus (point wise) $`X^g`$ consists of only (smooth) rational curves, and $`(4)`$ $`X^g`$ contains at least $`10`$ rational curves. Then $`(X,g)`$ is isomorphic to the pair $`(X_2,g_2)`$ up to isomorphisms. ###### Theorem 3.16. (Oguiso-Zhang , Theorem $`3`$). Let $`(X,g)`$ be a pair of a $`K3`$ surface $`X`$ and an automorphism $`g`$ of $`X`$ satisfying: $`(1)`$ $`g^3=id`$, $`(2)`$ $`g^{}\omega _X=\zeta _3\omega _X`$, $`(3)`$ The fixed locus (point wise) $`X^g`$ consists of only (smooth) rational curves and possibly some isolated points, and $`(4)`$ $`X^g`$ contains at least $`6`$ rational curves. Then $`(X,g)`$ is isomorphic to the pair $`(X_3,g_3)`$ up to isomorphisms. ###### Remark 3.17. (1) The $`K3`$ surfaces $`X_3`$ and $`X_2`$ in the theorems above have Picard number 20 and discriminants $`|det(\text{Pic}(X_3))|=3`$ and $`|det`$ $`(\text{Pic}(X_2))|=4`$. Indeed, $`X_3`$ (resp. $`X_2`$) is the only $`K3`$ surface with Picard number 20 and discriminant 3 (resp. 4). They are called the most algebraic $`K3`$ surfaces in Vinberg . For these two $`X=X_i`$, the infinite group $`\text{Aut}(X)`$ has been calculated by Vinberg. (2) A smooth $`K3`$ surface $`X`$ is called singular if the Picard number $`\rho (X)=20`$ (maximum). Such $`X`$ is uniquely determined by the transendental lattice $`T_X`$; see Shioda-Inose . Moreover, it has been proved there that a singular $`K3`$ surface has infinite group $`\text{Aut}(X)`$. See Oguiso’s theorem below for a stronger result. (3) Nikulin, Vinberg and Kondo have determined all $`K3`$ surfaces $`X`$ with finite $`\text{Aut}(X)`$; see Nikulin and Kondo . (4) In general, it is very difficult to calculate the full automorphism group $`\text{Aut}(X)`$ for a $`K3`$ surface $`X`$. This has been done by Keum and Kondo for a generic Jacobian Kummer surface and by Keum-Kondo for Kummer surfaces associated with products of some elliptic curves. (5) In their paper , Dolgachev and Keum have successfully calculated the full automorphism group $`\text{Aut}(X)`$ for general quartic Hessian surfaces, by embedding the rank $`16`$ Picard lattice into the rank $`26`$ lattice $`L`$ of signature $`(1,25)`$ (the sum of the rank $`24`$ Leech lattice and the standard hyperbolic plane). See also Kondo for the employment of such $`L`$ for Kummer surfaces. We end this section with the following striking result in (Oguiso ); see the same paper for more results in the new directions. ###### Theorem 3.18. (Oguiso ). Let $`X`$ be a singular $`K3`$ surface and let $`H:=Hilb^n(X)`$. Then each of $`\text{Aut}(X)`$, $`\text{Aut}(H)`$ and $`\text{Bir}(M)`$ contains the non-commutative free group $``$. ## 4. $`K3`$ groups of mixed type So far, we have considered $`K3`$ groups of purely symplectic or purely non-symplectic type. In this section, we shall consider $`K3`$ groups $`G`$ of the mixed type, i.e., the case where neither $`G_N`$ nor $`\mu _I`$ (with $`I=I(X,G)`$) in the basic exact sequence of the Introduction becomes trivial. On the one hand, $`|G_N||M_{20}|=960`$ by Theorem 2.2. On the other hand, $`|\mu _I|66`$ by Lemma 3.1. A finite $`K3`$ group $`G`$ (and the pair $`(X,G)`$) of the largest order has been determined by Kondo: ###### Theorem 4.1. (Kondo ). Let $`X`$ be a $`K3`$ surface and $`G\text{Aut}(X)`$ a finite subgroup. $`(1)`$ We have $`|G|4960`$. $`(2)`$ Suppose that $`|G|=4960`$. Then $`X`$ equals the Kummer surface $`Km(E_\sqrt{1}\times E_\sqrt{1})`$ where $`E_\sqrt{}1=/(+\sqrt{1})`$, and $`G=M_{20}.\mu _4`$. Moreover, the group $`G`$ and the action of $`G`$ on $`X`$ are unique up to isomorphisms. Besides the fact that $`M_{20}`$ is the finite symplectic $`K3`$ group of the largest order, it is also the the finite perfect $`K3`$ group of the largest order. Here are all of the finite perfect $`K3`$ groups (cf. Xiao , the list): $$M_{20},A_6,L_2(7)=\text{PSL}_2(7),A_5.$$ The first three groups are among the 11 maximum ones in Theorem . The last three groups above are also the only non-abelian finite simple $`K3`$ groups. The last two groups happen to be the two smallest (in terms of order) non-abelian finite simple groups. These four groups (together with a bigger $`\mu _I`$) determine the surface $`X`$ uniquely (see the theorems below). Let us start with an example of $`K3`$ group of mixed type. ###### Example 4.2. Let $`C_{168}=\{x_1x_2^3+x_2x_3^3+x_3x_1^3=0\}^2`$ be the Klein quartic curve of genus $`g=3`$. A well known theorem says that $`|\text{Aut}(C)|84(g1)`$ for any curve $`C`$ of genus $`g2`$. This Klein quartic attains the maximum $`|\text{Aut}(C_{168})|=84(31)=168=|L_2(7)|`$ (see Oguiso-Zhang for the action of $`L_2(7)`$ on $`C_{168}`$). Consider the quartic $`K3`$ surface $`X_{168}=\{x_0^4+x_1x_2^3+x_2x_3^3+x_3x_1^3=0\}^3`$. This is also the Galois $`/(4)`$-cover of $`^2`$ branched along the Klein quartic $`C_{168}`$. Clearly, there is a faithful $`L_2(7)\times \mu _4`$ action on $`X_{168}`$. We shall call the pair $`(X_{168},L_2(7)\times \mu _4)`$ the Klein-Mukai pair. ###### Theorem 4.3. (Oguiso-Zhang , Main Th, Prop. $`1`$). Let $`X`$ be a $`K3`$ surface such that $`G:=L_2(7)\text{Aut}(X)`$. Let $`\stackrel{~}{G}\text{Aut}(X)`$ be a finite subgroup containing $`G`$. $`(1)`$ $`G`$ is normal in $`\stackrel{~}{G}`$ with $`|\stackrel{~}{G}/G|4`$. $`(2)`$ Suppose that $`|\stackrel{~}{G}/G|=4`$. Then $`(X,\stackrel{~}{G})`$ is isomorphic to the Klein-Mukai pair $`(X_{168},L_2(7)\times \mu _4)`$. Next we turn to the another simple group $`A_6`$, also called the Valentiner’s group according to Dolgachev and Gizatullin (see the reference in Keum-Oguiso-Zhang ). It is a very important group and also the root of troubles in the theory of simple groups. Indeed, the $`A_6`$ is at the junction of the three simple groups series: recall that $`A_6=[M_{10},M_{10}]=\text{PSL}(2,9)`$. ###### Theorem 4.4. (Keum-Oguiso-Zhang , Th $`3.1`$, Prop $`4.1`$, Th $`5.1`$). $`(1)`$ Let $`F`$ be the unique $`K3`$ surface of Picard number $`20`$ whose transcendental lattice $`T_F`$ has the intersection matrix $`\text{diag}[6,6]`$ (see Shioda-Inose ). Then there is a group $`\stackrel{~}{A}_6:=A_6:\mu _4`$ (a split extension of $`A_6`$ by $`\mu _4`$) and a faithful action of it on $`F`$ (see Keum-Oguiso-Zhang Def $`2.7`$, Th $`3.1`$ or Th $`3.2`$ for the action and the group structure). $`(2)`$ Suppose that $`\stackrel{~}{G}`$ is a finite $`K3`$ group containing $`A_6`$. Then $`A_6`$ is normal in $`\stackrel{~}{G}`$, and the group $`\stackrel{~}{G}/A_6`$ is cyclic of order $`4`$. $`(3)`$ Suppose that $`(X,\stackrel{~}{G})`$ is a pair of a $`K3`$ surface and a subgroup $`\stackrel{~}{G}\text{Aut}(X)`$ containing $`A_6`$ as a subgroup of maximum index $`4`$. Then $`(X,\stackrel{~}{G})`$ is isomorphic to the pair $`(F,\stackrel{~}{A}_6)`$ in $`(1)`$. ###### Remark 4.5. The $`F`$ in the theorem above is also the minimal resolution of the hypersurface in $`^1\times ^2`$ (with coordinates $`([S:T],[X:Y:Z])`$) given by the equation (see Keum-Oguiso-Zhang Prop 3.5): $$S^2(X^3+Y^3+Z^3)3(S^2+T^2)XYZ=0.$$ The action of $`\stackrel{~}{A}_6`$ is still invisible from the equation (due to the trouble nature of $`A_6`$?) It is indeed constructed lattice-theoretically, where the Leech lattice and some deep results of S. Kondo and R. E. Borcherds are utilized. In the case of the smallest non-abelian simple group $`A_5`$, we have the complete classification of its finite $`K3`$ over-groups: ###### Theorem 4.6. (Zhang Th A, or Th E). Suppose that $`G`$ is a finite $`K3`$ group containing $`A_5`$ as a normal subgroup. Then $`G`$ equals one of the following (for their geometric realization, see Zhang Example $`1.10`$): $$A_5,S_5,A_5\times \mu _2,S_5\times \mu _2.$$ We now give an example of the largest (in terms of the order) finite solvable $`K3`$ group. ###### Example 4.7. Let $`X_4:=\{x_1^4+x_2^4+x_3^4+x_4^4=0\}^3`$ be the Fermat quartic $`K3`$ surface. Then $`(\mu _4)^4`$ acts diagonally on the coordinates while $`S_4`$ permutes the coordinates. So there is a natural action of $`\stackrel{~}{F}_{384}=(\mu _4^4:S_4)/\mu _4=(\mu _4^4/\mu _4):S_4`$ on $`X_4`$ (see Mukai ). This $`\stackrel{~}{F}_{384}`$ is a solvable group of order $`4^34!=2^93=3844`$. Note that the $`2`$-Sylow subgroup $`\stackrel{~}{F}_{128}`$ of $`\stackrel{~}{F}_{384}`$ is of course nilpotent and of order $`2^9=1284`$. Let $`G=\stackrel{~}{F}_{128}`$, or $`\stackrel{~}{F}_{384}`$, then in notation of the basic exact sequence in the Introduction, we have $`\mu _{I(X,G)}=\mu _4`$ and $`G_N=F_{128}`$, or $`F_{384}`$ respectively (as defined in Mukai ; note also that $`F_{384}`$ is also the second largest among the 11 maximum ones in Theorem 2.2). For both $`G`$, one can show that $`\text{Pic}(X_4)^G=H`$ and $`H^2=4`$ (see Oguiso , Th 1.2). Now we state a pretty result due to Oguiso: ###### Theorem 4.8. (Oguiso , Th $`1.2`$). $`(1)`$ Let $`X`$ be a $`K3`$ surface and $`G\text{Aut}(X)`$ a finite solvable subgroup. Then $`|G|2^93`$. Moreover, if $`|G|=2^93`$ then $`(X,G)`$ is isomorphic to the pair $`(X_4,\stackrel{~}{F}_{384})`$ (with the standard action) in the example above. $`(2)`$ Let $`X`$ be a $`K3`$ surface and $`G\text{Aut}(X)`$ a finite nilpotent subgroup. Then $`|G|2^9`$. Moreover, if $`|G|=2^9`$ then $`(X,G)`$ is isomorphic to the pair $`(X_4,\stackrel{~}{F}_{128})`$ (with the standard action) in the example above. ## 5. Log Enriques surfaces The concept of Log Enriques surface was first introduced in Zhang , Def 1.1. In the smooth case, they are just abelian surface (index 1), $`K3`$ surface (index 2), Enriques surface (index 2) and hyperelliptic surfaces (index $`2,3,4`$, or $`6`$). Log Enriques surface is closely related to the study of purely non-symplectic $`K3`$ groups. ###### Definition 5.1. A normal projective surface $`Y`$ with at worst quotient singularities is called a log Enriques surface if the irregularity $`h^1(Y,𝒪_Y)`$ $`=0`$ and if a positive multiple of the canonical divisor $`mK_Y0`$ (linearly equivalence). $`I=I(Y)=\mathrm{min}\{m_{>0}|mK_Y0\}`$ is called the index of $`Y`$. There is a Galois $`/(I)`$-cover $`\pi :X=Spec_{i=0}^{I1}𝒪_Y(iK_Y)Y`$ which is unramified outside non-Du Val singular locus of $`Y`$, i.e., outside the set of singular points which is not rational double. This $`X`$ is called the canonical cover of $`Y`$. One sees that $`g^{}\omega _X=\eta _I\omega _X`$, where $`\text{Gal}(X/Y)=g\mu _I`$ and $`\eta _I`$ is a primitive $`I`$-th root of $`1`$. ###### Remark 5.2. (1) In general, the canonical cover $`X`$ is a normal surface with at worst Du Val singularities and with $`K_X0`$. So either $`X`$ is a (smooth) abelian surface, or a $`K3`$ surface with at worst Du Val singularities. (2) Log Enriques surfaces are degenerate fibres of families of Kodaira dimension 0. Also the base space of elliptically fibred Calabi-Yau threefold $`\mathrm{\Phi }_D:ZY`$ with $`D.c_2(Z)=0`$ are necessarily log Enriques surfaces (see Oguiso ). (3) When the canonical cover $`X`$ of $`Y`$ is abelian, one has $`I(Y)=3`$, or $`5`$ and such $`Y`$ is unique up to isomorphisms; for examples or proof, see Zhang Example 4.2 and Theorem 4.1, or Blache , (1.2) and Theorem C. (4) In general, one has $`I(Y)21`$; see Blache Th C, and also Zhang Lemma 2.3. If $`I(Y)`$ is prime, then $`I(Y)19`$ (Lemma 3.1). Examples of prime $`I(Y)`$ are given in Zhang and Blache . When $`I(Y)`$ is prime, the classification of $`\text{Sing}(X)`$ for the canonical cover $`X`$ of $`Y`$, is done in Zhang , Main Th. (5) When the canonical cover $`X`$ of $`Y`$ is a normal $`K3`$, we let $`\stackrel{~}{X}X`$ be the minimal resolution. Then the determination of $`Y`$ is almost equivalent to that of $`(\stackrel{~}{X},\mu _I)`$, where $`\mu _I=\text{Gal}(X/Y)`$ acts purely non-symplectically on the (smooth) $`K3`$ surface $`\stackrel{~}{X}`$. Clearly, $`20\rho (\stackrel{~}{X})=r(Y)+\rho (X)r(Y)+1`$ with $`r(Y)`$ the number of components in the exceptional divisor of the resolution $`\stackrel{~}{X}X`$. Here $`\rho (Z)`$ denotes the Picard number. Hence $`r19`$. (6) For rational log Enriques surfaces $`Y`$ with maximum $`r(Y)=19`$, the type $`\text{Sing}(X)`$ has been determined. They are one of 7 Dynkin types: $$D_{19},D_{16}+A_3,D_{13}+A_6,D_7+A_{12},D_7+D_{12},D_4+A_{15},A_{19}.$$ For each of the 7 Dynkin types above, there is exactly one (up to isomorphisms) rational log Enriques surface $`Y`$ such that the singular locus $`\text{Sing}(X)`$ of the canonical cover $`X`$ of $`Y`$ has that Dynkin type; see Oguiso-Zhang Th 1, Th 2 and Main Th. Start with a log Enriques surface $`Y`$, by passing to a maximum crepant partial resolution (i.e., resolving Du Val singularities), we may assume the maximality condition for $`Y`$: (\*) Any birational morphism $`Y^{}Y`$ from another log Enriques surface $`Y^{}`$ must be an isomorphism. We now state a uniqueness result for the three largest prime indices: ###### Theorem 5.3. (Oguiso-Zhang , Cor. 4). Let $`Y`$ be a log Enriques surface with $`I=I(Y)=19,17`$ or $`13`$ and satisfying the maximality condition (\*) above. $`(1)`$ We have $`YX_I/g_I`$ when $`I=19`$ or $`17`$. $`(2)`$ We have $`Y\overline{X}_{13}/g_{13}`$ when $`I=13`$. Here the pairs $`(X,g_I)`$ are given in Example 3.7, and $`\overline{X}_{13}`$ is obtained from $`X_{13}`$ by contracting the unique rational curve in the fixed locus (point wise) $`X_{13}^{g_{13}}`$. (see Zhang Ex $`5.75.8`$, Ex $`7.3`$ for different constructions). We end the paper with the final remark. ###### Remark 5.4. (1) In their paper , Dolgachev-Keum have extended Mukai’s classification Theorem 2.2 to characteristic $`p`$ case. For smaller $`p`$, some new groups of automorphisms (not in Mukai’s list) appear, see ibid and Dolgachev-Kondo . (2) In his paper , Theorem 1.5, Oguiso considers the subtle behavior of $`\text{Aut}(X_t)`$ for $`K3`$ surfaces in a 1-dimensional family. On the one hand, for generic $`t`$ the group $`\text{Aut}(X_t)`$ contains a fixed subgroup with bounded index. On the other hand, he also produces an example where $`\text{Aut}(X_t)`$ is infinite for generic $`t`$ but is finite for some special $`t`$. (3) Many interesting results could not be included in the paper due to the limitation of the knowledge of the author and the constraint of the space.
warning/0506/astro-ph0506511.html
ar5iv
text
# The Spectral Evolution of Transient Anomalous X-ray Pulsar XTE J1810–197 ## 1. Introduction The bright 5.54 s X-ray pulsar XTE J1810$``$197 is the second example of a Transient Anomalous X-ray Pulsar (TAXP) and the first one confirmed by measuring a rapid spin-down rate. All of its observed and derived physical parameters are consistent with classification as an Anomalous X-ray Pulsar (AXP), one that had an impulsive outburst sometime between 2002 November and 2003 January, when it was discovered serendipitously by Ibrahim et al. (2004) using the Rossi X-ray Timing Explorer (RXTE). Its flux was observed to be declining with an exponential time constant of $`269\pm 25`$ days from a maximum of $`F_X(210\mathrm{keV})6\times 10^{11}`$ ergs cm<sup>-2</sup> s<sup>-1</sup>. The source was then localized precisely using two Target of Opportunity (ToO) observations with the Chandra X-ray Observatory by Gotthelf et al. (2004, hereafter Paper I) and Israel et al. (2004). In comparison, archival detections by several X-ray satellites indicate a long-lived quiescent baseline flux of $`F_X(0.510\mathrm{keV})7\times 10^{13}`$ ergs cm<sup>-2</sup> s<sup>-1</sup> lasting at least 13 years and possibly for 23 years prior to the outburst (Paper I). Fading of an IR source within the Chandra error circle, similar to ones associated with other AXPs, confirmed its identification with XTE J1810$``$197 (Rea et al., 2004a, b). The X-ray spectra and pulse profiles from three observations obtained with XMM-Newton during the decline of the outburst were studied by Halpern & Gotthelf (2005, hereafter Paper II). The short duty cycle of activity of XTE J1810$``$197 suggests the existence of a significant population of as-yet unrecognized, although not necessarily undetected, young neutron stars (NSs). In this paper we present a new XMM-Newton observation of XTE J1810$``$197. This data set, combined with previous XMM-Newton observations acquired over the last year, allows us to characterize the spectral evolution of a Transient AXP. We show that the decay rates of the two fitted X-ray spectral components are distinct, and can be extrapolated in time to approach the previous quiescent spectrum. For consistency with Papers I and II, we express results in terms of a maximum distance of $`d=5`$ kpc to the pulsar. However, there is reason to believe that XTE J1810$``$197 is significantly closer than this, and we take into account a more realistic estimate of 2.5 kpc in our discussion of proposed physical models for the X-ray emission and outburst mechanism. ## 2. Observations A fourth XMM-Newton observation of XTE J1810$``$197 was obtained on 2004 September 18. The previous three observations are described in Paper II. We use the data collected with the European Photon Imaging Camera (EPIC, Turner et al. 2003) which consist of three CCD imagers, the EPIC pn and the two EPIC MOSs, each sensitive to X-rays in the $`0.112`$ keV energy range. In the following we concentrate on data taken with the EPIC pn detector, which provided a timing resolution of 48 ms in “large window” mode, for ease of comparison with the earlier data sets. The fast readout of this instrument ensures that its spectrum is not affected by photon pileup. Data collected with the two EPIC MOS sensors used the “small window” mode and “timing” mode, providing a time resolution of 0.3 s and 1.5 ms, respectively. For the new data set, we followed the reduction and analysis procedures used for the previous XMM-Newton observations of XTE J1810$``$197, as outlined in Paper II. The new observation is mostly uncontaminated by flare events and the filtered data set resulted in a total of 26.5 ks of good EPIC pn exposure time (24.4 ks live time). We checked for timing anomalies that were evident in some previous EPIC pn data sets and found none. Photon arrival times were converted to the solar system barycenter using the Chandra derived source coordinates R.A. $`18^\mathrm{h}09^\mathrm{m}51.^\mathrm{s}08`$, decl. $`19^{}43^{}51.^{\prime \prime }7`$ (J2000.0) given in Paper I. ### 2.1. Spin-down Evolution of XTE J1810$``$197 The barycentric pulse period of XTE J1810$``$197 measured at each XMM-Newton observing epoch is given in Table 1. These are derived from photons obtained with both the EPIC pn and MOS cameras, with the exception of the short 2003 October 12 observation for which only EPIC pn data of sufficient time resolution is available. The errors are the 95% confidence level determined from the $`Z_1^2`$ test. As shown in Figure 1, the four period measurements can be fitted to yield a mean spin-down rate of $`\dot{P}=(8.1\pm 0.7)\times 10^{12}`$ s s<sup>-1</sup> over the year-long interval. This implies a characteristic age $`\tau _\mathrm{c}10,800`$ yr, surface magnetic field $`B_\mathrm{s}2.1\times 10^{14}`$ G, and spin-down power $`\dot{E}1.9\times 10^{33}`$ erg s<sup>-1</sup>, comparable to the earlier values (Ibrahim et al., 2004, and Paper II). Deviations from a constant $`\dot{P}`$ are evident, however, since Ibrahim et al. (2004) fitted values in the range $`(1.12.2)\times 10^{11}`$ s s<sup>-1</sup> in the first 9 months of the outburst. As discussed in paper II (§2.2), timing glitches in AXPs can result in large increases in their period derivatives \[$`\mathrm{\Delta }\dot{P}/\dot{P}1`$; e.g., 1E 2259+586 Iwasawa, Koyama, & Halpern (1992); Kaspi et al. (2003) and 1RXS J170849.0–400910 Dall’Osso et al. (2003)\]. Accordingly, it is possible that XTE J1810$``$197 experienced a glitch and an increase in $`\dot{P}`$ at the time of its outburst, and that $`\dot{P}`$ is now relaxing to its long-term value. But since there is no prior ephemeris for XTE J1810$``$197 in its quiescent state, we do not know if its outburst was triggered by a glitch. Alternatively, the spin-down torque could have been enhanced in the early stages of the outburst by an increase over the dipole value of the magnetic field strength at the speed-of-light cylinder (Thompson, Lyutikov, & Kulkarni, 2002), or by a particle wind and Alfvén waves (Thompson & Blaes, 1998; Harding, Contopoulos, & Kazanas, 1999), effects that are expected to decline after the first few months. ### 2.2. Spectral Analysis and Results XMM-Newton observations of XTE J1810$``$197 have shown that its spectrum is equally well fitted by a power-law plus blackbody model, as commonly quoted for AXPs, or a two-temperature blackbody model. In Paper II we argued that the two-temperature model is more physically motivated, while the power-law plus blackbody model suffers from physical inconsistencies. As we shall show, the new data bolster these arguments, so we concentrate mainly on the double blackbody model in this work. Table 1 presents a summary of spectral results from all four XMM-Newton observations of XTE J1810$``$197. As with the earlier data, the 2004 September source spectrum was accumulated in a $`45^{\prime \prime }`$ radius aperture which encloses $`95\%`$ of the energy. Background was taken from a circle of the same size displaced $`2\stackrel{}{\mathrm{.}}3`$ along the readout direction. The spectra were grouped into bins containing a minimum of 400 counts (including background) and fitted using the XSPEC package. For this analysis the column density is held fixed at $`N_\mathrm{H}=6.5\times 10^{21}`$ cm<sup>-2</sup>, the value determined from previous fits, which are all consistent. The best fit to the two-temperature blackbody model yields temperatures of $`kT_1=0.25\pm 0.01`$ keV and $`kT_2=0.67\pm 0.01`$ keV with a fit statistic of $`\chi _\nu ^2=1.2`$ for 188 degrees of freedom (Fig. 2). Each of these two temperatures, which we refer to as “warm” and “hot”, respectively, remained essentially the same, nominally to within $`4\%`$ of that reported for the first XMM-Newton observation (see Table 1). However, the warm and hot blackbody luminosities declined by 33% and 70%, respectively, in one year. With the temperature of the hot component remaining essentially constant, its decline in luminosity is attributable to a decrease in its emitting area, $`A_2`$. The most recent value, $`A_2=2.1\times 10^{11}d_5^2`$ cm<sup>2</sup>, is $`1\%`$ of the NS surface area. In the case of the warm component, the decay in flux is modest, so that it is not yet possible to decide within the errors whether the temperature or the area is the primary variable. The most recent area measurement of the warm component, $`A_1=9.1\times 10^{12}d_5^2`$ cm<sup>2</sup>, is $`50\%`$ of the NS surface area. Phase-resolved spectroscopy using the new XMM-Newton data set is compared here to that reported in Paper II. As before, we fit for the intensity normalization in each of 10 phase bins with the two temperatures and column density fixed at the phase-averaged value listed in Table 1. This is equivalent to determining the relative projected area of emission as a function of rotation phase. The fits are again found to be each statistically acceptable as a set, so no significant test can be made for variations in additional parameters such as the temperatures or column density. The results of the phase-resolved spectral fits are shown in Figure 3. Comparison between the data sets taken a year apart shows that the modulation with phase of each blackbody component has remained steady to within the measurement errors. The phase alignment of the two temperature components has remained the same, and the pulses peak at the same phase. This is consistent with the picture of a small, hot region surrounded by a warm, concentric annulus that occupies $`1/2`$ the surface area of the star. Neither component disappears at any rotation phase. In particular, if the hot component were completely eclipsed, the spectral decomposition in Figure 2 indicates that the light curves at $`E>3`$ keV should dip to zero; clearly they do not. ### 2.3. Long-term Flux Decay With the set of XMM-Newton measurements spanning a year, a more accurate characterization of the X-ray flux decay of XTE J1810$``$197 is possible than that reported in the initial RXTE study of Ibrahim et al. (2004). The bolometric flux over time, shown in Figure 4, reveals a new and complex behavior. While the RXTE data were consistent with an exponential decay of time constant $`\tau =269\pm 25`$ days, the subsequent XMM-Newton measurements show that the inferred bolometric luminosity of the two spectral components are not declining at the same rate. From the data presented in Table 1 we find that the luminosity of the hotter temperature component falls exponentially with $`\tau _2=300`$ days, while the warm component decreases with $`\tau _1=900`$ days, which, although less reliably characterized, is clearly longer than $`\tau _2`$. The shorter time constant is very close to the one describing the RXTE data. Furthermore, the RXTE spectrum itself is fitted best by a single temperature blackbody of $`kT=0.7`$ keV (Roberts et al., 2004), without a power-law contribution, which is consistent with our interpretation of the XMM-Newton spectrum. Since RXTE is sensitive only at energies $`>2`$ keV where the hotter blackbody component dominates, the overall X-ray decay from the beginning of the outburst to the present appears consistent with separate exponential time constants corresponding to the two distinct thermal-spectrum components. We also note that it is not possible to fit an alternative power-law temporal decay to the hot blackbody flux; such a model would require a decay index that steepens with time. As shown in Figure 4, we expect that the X-ray flux of XTE J1810$``$197 will return to its historic quiescent level by the year 2007. If the quiescent spectrum is to match that observed before the current outburst, then it would likely resemble the ROSAT observation of 1992 March 7. Although of relatively poor quality, the ROSAT spectrum can be reasonably well fitted with a single blackbody of $`kT=0.18\pm 0.02`$ keV covering $`1.2\times 10^{13}d_5^2`$ cm<sup>2</sup> and $`L_{BB}(\mathrm{bol})=1.3\times 10^{34}d_5^2`$ ergs s<sup>-1</sup> (Paper I). This blackbody is significantly cooler and larger than the outburst warm component of $`kT_1=0.25`$ keV and $`A_1=9.1\times 10^{12}d_5^2`$ cm<sup>2</sup>, so it should be present even now, although masked by the fading outburst emission. We estimate that the $`kT=0.18`$ keV blackbody spectrum measured with ROSAT would contribute 30% of the latest (2004 September) XMM-Newton measured flux at 0.5 keV. Therefore, we expect that such a cooler component will begin to dominate the soft X-ray spectrum in late 2006, consistent with the projection of the flux in Figure 4. ### 2.4. Pulse Shape Evolution The energy-resolved folded light curves from the 2003 September 8 XMM-Newton observation (Fig. 5a) show that the pulse peak, in general, is somewhat narrower than a sinusoid, an effect that is more pronounced at higher energy. Examination of the light curves measured one year later (2004 September 18; Fig. 5d) shows a clear energy-dependent change. This is easily seen by forming the ratio of the curves taken at the two epochs, normalized by their count-weighted mean. These ratio curves, presented in Fig. 6, show changes in amplitude of up to 20%. Specifically, the pulse shape of the lower-energy light curves has evolved to a smaller amplitude, more sinusoidal profile. This change is highly significant as determined by the reduced $`\chi _\nu ^2`$ (24 degrees-of-freedom) for the null hypothesis of a flat line (see labels on Fig. 6). In particular, the most pronounced variation is found in the 1.0–1.5 keV band with $`\chi _\nu ^2=6.3`$ corresponding to an infinitesimal probability of the pulse shape being constant. In contrast, the pulse shape of the higher energy bins remained essentially unchanged, as evidence by their $`\chi _\nu ^2`$ values near unity. This trend is reflected in the two intervening observations (Fig. 5c and 5d) as well. The pulse profile must be made up of some combination of pulsed emission intrinsic to each of the two measured spectral components. In any given energy band, the changing relative contribution of the soft and hard spectral components forces the combined pulse shapes to evolve in time. One way of quantifying this change is to define the pulsed fraction $`f_p`$, the fraction of flux above the minimum in the light curve. This is found to increase smoothly with energy from 34% at less than 1 keV to $`51\%`$ above 5 keV at the earlier epoch (see Table 2 and Fig. 5). At the latest epoch reported herein, the pulsed fraction in the lower energy band decreased to 26% while it remained essentially unchanged at the higher energy. The simplest hypothesis for modeling the light curve evolution is to assume that the intrinsic light curve of each spectral component is steady in time. If so, the pulsed fraction of the light curve at a given energy and epoch $`f_p(E,t)`$ can be expressed as a linear combination of two light curves of fixed pulsed fractions $`f_{BB1}`$ and $`f_{BB2}`$ weighted by the ratio of counts from the respective spectral component $`N_{BB1}(E,t)`$ and $`N_{BB2}(E,t)`$, $`f_p(E,t)=`$ $`{\displaystyle \frac{f_{BB1}N_{BB1}(E,t)+f_{BB2}N_{BB2}(E,t)}{N_{BB1}(E,t)+N_{BB2}(E,t)}}`$ (1) $`=`$ $`f_{BB1}[1R(E,t)]+f_{BB2}R(E,t),`$ (3) where $`R(E,t)=N_{BB2}(E,t)/[N_{BB1}(E,t)+N_{BB2}(E,t)]`$ is the normalized flux ratio. This expression is only valid in the case that the minima of the two light curves coincide, as is evident for the XTE J1810$``$197 profiles. For an exponential decay of the component luminosities, the time-dependent model for the pulsed fraction is simply, $$f_p(E,t)=\frac{f_{BB1}+f_{BB2}r(E,t_o)e^{(tt_o)/\tau _d}}{1+r(E,t_o)e^{(tt_o)/\tau _d}},$$ (4) where $`r(E,t_o)=N_{BB2}(E,t_o)/N_{BB1}(E,t_o)`$ is the flux ratio at time $`t_o`$ and $`\tau _d=(1/\tau _21/\tau _1)^1=450`$ d is the differential time constant. We can test this hypothesis directly by fitting for $`f_{BB1}`$ and $`f_{BB2}`$ using a joint least-squares fit to the set of 24 (six energy bands at four epochs) measured pulsed fractions $`f_p(E,t)`$ given in Table 2 and the ratio $`R(E,t)`$ for the double blackbody spectral model tabulated in Table 3. Treating the ratio $`R(E,t)`$ as the independent variable, the best fit yields $`f_{BB2}=52\pm 3\%`$ for the hotter component and $`f_{BB1}=26\pm 2\%`$ for the cooler one. The fit statistic is $`\chi _\nu ^2=1.17`$ for 22 degrees-of-freedom, corresponding to a probability $`\mathrm{}(1.17)=0.27`$. The lower panel of Fig. 4 shows the best fit model for the decay of the the 1.0–1.5 keV pulsed fraction over time and its extrapolation to an epoch dominated by the warm spectral component. On the other hand, if we apply the blackbody plus power-law model, the corresponding values of $`R(E,t)`$, also listed in Table 3, are completely different and the result is $`\chi _\nu ^2=1.70`$ for 22 degrees-of-freedom, corresponding to a probability $`\mathrm{}(1.70)=0.021`$, 13 times less likely relative to the double blackbody model, for the assumption of fixed intrinsic pulsed fractions. Evidently the pulsed fraction of the hotter component is much higher than for the cooler component. This difference would account for the gradual shift from a sharper, more triangular pulse shape found at higher energies, to a rounder and more symmetric sinusoidal light curve seen at lower energies. As was hypothesized above, the decrease in pulsed fraction with time at low energies follows from the fact that the hot spectral component is decaying more rapidly than the cooler one, and so its contribution to the pulsed fraction in the low-energy bin is declining. To the extent that we are not able to discern differences greater than $`4\%`$, which represents the measurement uncertainty, the pulsed fractions $`f_{BB1}`$ and $`f_{BB2}`$ are individually independent of energy. However, we also cannot rule out a model in which $`f_{BB1}`$ itself increases with increasing energy. It is useful to model the shape of the light curves of XTE J1810$``$197 in detail to further quantify their evolution. Overall, the pulse profiles suggest some linear combination of sinusoidal and triangular functions, for the soft and hard X-rays, respectively. In the following, we model the photon counts as a function of phase, $`N(\varphi )`$, at a given energy $`E`$ and epoch $`t`$ by the two-component model, $$N(\varphi ;E,t)=N_S(\varphi ;E,t)+N_T(\varphi ;E,t),$$ where $$N_S(\varphi ;E,t)=\alpha (E,t)[\mathrm{\hspace{0.17em}1}+\mathrm{cos}(\varphi \varphi _S)]+\gamma _S(E,t)$$ and $$N_T(\varphi ;E,t)=\{\begin{array}{cc}\beta (E,t)[\mathrm{\hspace{0.17em}1}2|\varphi \varphi _T|/\delta (E,t)]+\gamma _T(E,t),\hfill & \\ if|\varphi \varphi _T|<\delta /2\hfill & \\ \gamma _T(E,t),if|\varphi \varphi _T|\delta /2\hfill & \end{array}$$ (5) Here, $`\alpha `$ is the amplitude of the pulsed signal and $`\gamma _S`$ represent the “unpulsed,” or minimum level, for the sinusoidal component, $`\beta `$ and $`\gamma _T`$ are the analogous parameters for the triangular component, and $`\delta `$ is the duty cycle (full width) of the triangular pulse. Initial fits to the light curves indicate that the model can be constrained by fixing the triangular width $`\delta `$ to a single value for all observations. The same proves true for the relative phase $`\varphi _S\varphi _T`$ between the peak fluxes of the two components. Accordingly, we used a bootstrap approach to determine these shape parameters. After finding a global fit by iteration, the triangular width is set at $`\delta /2\pi =0.53`$ cycles. We then determined $`\varphi _S\varphi _T`$, which is found to be consistent with zero. Finally, the absolute phase $`\varphi _S`$ at each epoch was fixed. We therefore conclude that there is no need to allow energy or time dependence for the individual component pulse shapes. With the shape of the pulse components fixed, we fit the restricted model with just $`\alpha ,\beta `$, and $`\gamma `$ ($`=\gamma _S+\gamma _T`$, since the $`\gamma `$’s are indistinguishable in this fit) as free parameters. This model fitted to all 24 individual pulse profiles yields a reduced $`\chi _\nu ^2`$ ranging from $`0.62.0`$ with a typical value of $`\chi _\nu ^2=1.2`$, corresponding to a probability of $`\mathrm{}(\chi _\nu ^2)=0.23`$ for 22 degrees-of-freedom. The best-fit model parameters, derived pulsed fractions, and fit statistic for each light curve are presented in Table 2. A clear trend is seen as the pulsed fraction decreases with time. The best-fit models are shown in Figure 5 overlaid on plots of the energy-resolved pulsed profiles for the four epochs. We checked that a single component model, either a pure sinusoidal or triangular pulse, does not adequately characterize the data based on an unacceptably large $`\chi _\nu ^2`$ for many of the fits. It is possible to further reduce the number of free parameters in the fit by constraining the coefficients $`\alpha ,\beta `$, and $`\gamma `$ so that the phase-averaged photon count ratios $`N_S(E,t)/N_T(E,t)`$ are identical with the ratios derived from the two-temperature blackbody spectral fits in each energy interval at each epoch. That is, we test the hypothesis that the warm blackbody is responsible for only the sinusoidal pulse and the hot blackbody for only the triangular pulse by requiring $`N_S(E,t)/N_T(E,t)=N_{BB1}(E,t)/N_{BB2}(E,t)`$ at all epochs. We find that, while reasonable fits to the light curves are possible for the lower and highest energy bands, fits in the $`1.55.0`$ keV range proved unacceptable. This argues against a simple one-to-one correspondence between the spectral and the chosen light curve component decomposition. It is possible that we have not started with the correct spectral or light curve basis pairs or that the basic pulse profiles are each energy dependent. It is also possible that there may be a third, unfitted softer component associated with the quiescent flux, as seen by ROSAT (see §2.3). Continuing observation of the spectrum as the source fades should clarify the relationship between spectral and light curve components. ## 3. Interpretation ### 3.1. Spectral Evolution: Constraining Models While the $`0.58`$ keV X-ray spectra of AXPs (including XTE J1810$``$197) clearly require a two-component fit, it is not possible to prove from spectra alone that either of two very different models, namely a power-law plus blackbody, or a two-temperature blackbody, is correct. For XTE J1810$``$197, either model yields an acceptable chi-square when fitted to the X-ray spectrum. However, in Paper II we presented physical arguments against the properties of the particular power law that results from the power-law plus blackbody fit. Unlike some rotation-powered pulsars whose spectra are fitted by hard power laws plus soft blackbodies, in AXPs the roles of these two components are reversed. For XTE J1810$``$197, most of the X-rays belonging to the $`\mathrm{\Gamma }3.7`$ power-law component are lower in energy than those fitted by the blackbody (see Figure 5a of Paper II). This steep power law must turn down sharply just below the X-ray band in order not to exceed the faint, unrelated IR fluxes. As shown in Paper II, this would be difficult to achieve in a synchrotron model. And in a Comptonization model, where are the seed photons? Even if such a sharp cutoff existed at the low end of the EPIC energy band, a model in which the power-law component is due to Compton upscattering of thermal X-rays from the surface does not explain why most of the power-law photons have lower energy than the observed blackbody photons unless there is a larger source of seed photons at energies below 0.5 keV. However, such a seed source would have to be thermal emission from the neutron star. Since the two-temperature model uses a large fraction of the surface area for the $`0.52`$ keV photons, we prefer to regard the X-rays in this band as plausible seeds for, rather than as the product of, inverse Compton scattering. Supporting this assumption in the specific case of XTE J1810$``$197 are the shapes of its pulse profiles. The smooth increase in pulsed fraction with energy, and the strict phase alignment at all energies, are unnatural under the hypothesis of two different emission mechanisms and locations. The evolving spectral shape and pulse profiles of XTE J1810$``$197 during its decline from outburst offer new and complementary evidence concerning the appropriate spectral decomposition. We find that the the pulsed fraction is declining with time at low energies, while remaining essentially constant at high energies. This fact is consistent with the changing contributions of fluxes in the two-temperature fit as the two components decline at different rates. The warm and hot components contribute significantly to the soft X-rays, whereas the hard X-rays are supplied entirely by the hot component. This is quantified in Table 3, where it is seen that the hot blackbody component accounts for 30–60% of the flux below 1.5 keV, and $`>99\%`$ of the flux above 3 keV at all times. Under this spectral decomposition, the pulsed fraction at energies above 3 keV is large and unchanging because one and only one spectral component is present at those energies, and it has the larger pulsed fraction of the two spectral components. In contrast, the pulsed fraction at low energies ($`<1.5`$ keV) is understood to decrease in time because the hot component makes only a fractional contribution to its light curve, and its fraction decreases faster than the warm spectral component, which has the intrinsically smaller pulsed fraction. This effect is also illustrated in Figure 4, where the modeled pulse fraction is compared to the data. In contrast, consider how the spectral components would contribute to the various energy bands in the power-law plus blackbody spectral model. Table 3 shows that, unlike the two blackbody model, a fitted power-law component accounts for 79% of the soft ($`<1.5`$ keV) X-rays at all epochs, while decreasing its contribution to the hard ($`>3`$ keV) X-rays over time. In this model, the pulsed fraction should remain constant at low energies because only the power law is present there, while at high energies the pulsed fraction should vary because both the blackbody and the power-law components are present, in varying proportions. This spectral decomposition would drive an evolution of the pulse shapes that is opposite of what is observed. Thus, we find that the detailed evolution of the X-ray emission from XTE J1810$``$197 further supports the assumption of a purely thermal surface emission model, and leaves no evidence of a steep power law in the $`0.58`$ keV band, as is commonly fitted to individual observations of AXPs. Our decomposition of the pulsed light curves is also consistent with the absence of modulation in the ROSAT observation of 1992 March 7, to an observed upper limit of 24% in the $`0.22.0`$ keV PSPC energy band. Since the ROSAT spectrum is fitted by an even cooler blackbody of $`kT=0.18`$ keV, it can be expected that the pulsed fraction in quiescence will fall below the 25% value fitted to the warm blackbody in XMM-Newton data once that component fades away. ### 3.2. Re-evaluation of the Distance to XTE J1810$``$197 For consistency with Papers I and II, we have parameterized results here in terms of a distance of 5 kpc. However, this was considered an upper limit based on an H I absorption kinematic distance to the neighboring supernova remnant G11.2$``$0.3 (Green et al., 1988) and its X-ray measured column density $`N_\mathrm{H}1.4\times 10^{22}`$ cm<sup>-2</sup> (Vasisht et al., 1996). In comparison, the column density fitted to a power-law plus blackbody model for XTE J1810$``$197 was $`N_\mathrm{H}=1.02\times 10^{22}`$ cm<sup>-2</sup>. Since on physical grounds we strongly prefer the double blackbody model, which requires $`N_\mathrm{H}=0.65\times 10^{22}`$ cm<sup>-2</sup>, it seems appropriate to reduce the distance estimate to one that is more compatible with the smaller $`N_\mathrm{H}`$ value. We now assume that the distance to XTE J1810$``$197 is 2.5 kpc, and explore the consequences of this revision. The effect is to reduce the inferred X-ray luminosities and blackbody surface areas. For $`d=2.5`$ kpc, the surface area emitting the warm blackbody component is $`A_1<3\times 10^{12}`$ cm<sup>-2</sup> at all times, which is less than 1/6 the area of a neutron star. This makes it easier to understand how the pulsed fraction of the softest X-rays can be as large as $`26\%`$. We can also revise the estimate of the total energy emitted in the outburst. At a distance of 2.5 kpc, the bolometric luminosity quoted in Paper II is $`L_{\mathrm{BB2}}=1.5\times 10^{35}d_{2.5}^2e^{(tt_0)/300\mathrm{days}}`$ ergs s<sup>-1</sup> with respect to the initial time $`t_0`$ of the outburst observed by RXTE (Ibrahim et al., 2004), which corresponds to an extrapolated energy of $`3.9\times 10^{42}d_{2.5}^2`$ ergs. Since RXTE is only sensitive to the hotter blackbody component, we should add another contribution from the XMM-Newton measured warm component, which can be estimated as $`L_{\mathrm{BB1}}=1.7\times 10^{34}d_{2.5}^2e^{(tt_0)/900\mathrm{days}}`$ ergs s<sup>-1</sup>. The extrapolation required to integrate the contribution of the warm component is less certain than for the hot component, but it corresponds to an energy of only $`1.3\times 10^{42}d_{2.5}^2`$ ergs. The total estimated energy is then $`5.2\times 10^{42}d_{2.5}^2`$ ergs, which is comparable to the amount of heat assumed to be deposited in the crust during a deep heating event (Lyubarsky, Eichler, & Thompson, 2002), or the extra energy stored in an azimuthally twisted external magnetic field (Thompson et al., 2002). In the following section, we discuss how the observations might probe the actual mechanism of the outburst. ### 3.3. Heating or Cooling? Two mechanisms have been invoked to explain the late-time “afterglow” of an outburst of a magnetar. The first involves surface heating by long-lived currents flowing on closed but azimuthally twisted external magnetic field lines, and Comptonization of the resulting surface X-ray emission by the same particles (Thompson et al., 2002). The originating event could be a sudden fracture in the crust in response to twisting of the internal magnetic field, or a plastic deformation of the crust that gradually transfers internal magnetic twist to the external field. It is thought that the major outbursts of SGRs are magnetically trapped fireballs triggered by a sudden fracture, but this is not known to apply to XTE J1810$``$197 because no bursting behavior was seen, and no prior rotational ephemeris exists to test for a glitch. In any case, the extra energy above that of a pure dipole, $`\mathrm{\Delta }E`$, stored in a twisted magnetic field external to the star is, $$\mathrm{\Delta }E=1.4\times 10^{44}\mathrm{\Delta }\varphi _{\mathrm{N}\mathrm{S}}^2\left(\frac{B_\mathrm{p}}{10^{14}\mathrm{G}}\right)^2\left(\frac{R_{\mathrm{NS}}}{10\mathrm{km}}\right)^3\mathrm{ergs},$$ (6) where $`\mathrm{\Delta }\varphi _{\mathrm{N}\mathrm{S}}<1`$ rad is the azimuthal twist from north to south hemisphere (Thompson et al., 2002). This is enough to account for the extrapolated $`5.2\times 10^{42}d_{2.5}^2`$ ergs radiated if $`\mathrm{\Delta }\varphi _{\mathrm{N}\mathrm{S}}`$ is a small fraction of a radian. In this model, the X-ray spectrum is a combination of surface thermal emission, and Compton scattering and cyclotron resonant scattering of the thermal X-rays. Compton upscattered flux may result in power-law emission that dominates at high energies such as that reported above 10 keV for the AXP in Kes 73 (Kuiper et al., 2004). However, the flat power-law index ($`\mathrm{\Gamma }0.94`$) does not contribute to the softer spectrum, in particular below 2 keV. So we consider that the potential power-law tail in the fainter XTE J1810$``$197 may no be detectable yet. The decay timescale is determined by the rate at which power is consumed by the magnetospheric currents, which is comparable to that needed to power the observed X-ray luminosities of AXPs in general and XTE J1810$``$197 in outburst. Furthermore, it is the evolution of surface heating that basically determines the instantaneous X-ray luminosity and its decay, but more specific predictions of decay curves have not been made. In the second picture, a deep crustal heating event, heat deposited suddenly by rearrangement of the magnetic field is gradually transferred by diffusion and radiation, and the radiated X-rays are purely thermal. While the originating event might have been a fracture of the crust as in the first scenario, an assumption that the heat is evenly distributed throughout the crust results in a particular decay curve from one-dimensional models of heat transfer (Lyubarsky et al., 2002). In this model, the heating is virtually instantaneous, within $`10^4`$ s, while it is the rate of conduction cooling that determines the observed decay curve. Lyubarsky et al. (2002) assumed a deposition of $`10^{25}`$ ergs cm<sup>-3</sup> to a depth of 500 m, which is $`1\%`$ of the magnetic energy density at the surface. If occurring over the entire crust, this is $`1\times 10^{43}`$ ergs, comparable to the total X-ray energy emitted during the decay of XTE J1810$``$197. However, Lyubarsky et al. (2002) showed that the resulting cooling luminosity would follow a $`t^{0.7}`$ power law, which is rather slow compared to the observed exponential decay of the hot component of XTE J1810$``$197, while 80% of the energy should be conducted into the neutron star core rather than radiated from the surface. It is not obvious if the behavior of XTE J1810$``$197 supports or excludes either of these models. One interesting possibility is that the slowly decaying warm component is the radiation from a deep heating event that affected a large fraction of the crust, while the hot component of the spectrum is powered by external surface heating at the foot-points of twisted magnetic field lines by magnetospheric currents that are decaying more rapidly. It is not yet clear if the flux decay of the warm component is primarily an effect of decreasing temperature or decreasing area, because the decay has so far been modest. It is possible that the decay of the warm component is consistent with a $`t^{0.7}`$ power law rather than an exponential of $`\tau _1=900`$ days. Fits to $`L_{BB1}`$ as a function of time allows a power-law decay index in the range $`0.3`$ to $`0.7`$. In contrast, because of its more rapid decay, it is evident that the hotter component is declining exponentially, in area rather than in temperature. This shrinking in area might be easier to understand as a decay of the currents or a rearrangement of the magnetic field lines that are channeling the heating on the surface. But there is no good evidence yet of an inverse Compton scattered component from the particles responsible for that current, at least not at energies $`<8`$ keV. Perhaps the solid angle subtended by the twisted field lines in the magnetosphere is too small to scatter most of the thermal photons. Each of the models is challenged in some aspect by the observations, but each may still have some applicability. Observing the decay through to quiescence should help to clarify the situation. ### 3.4. Constraints on Emission and Viewing Geometry The pulsed light curves offer an additional diagnostic of the emission and viewing geometry. If we adopt the hypothesis that virtually all of the $`0.58`$ keV flux is surface thermal emission, then the symmetry of the light curves and their strict pulse-phase alignment as a function of energy argue for a concentric geometry in which a small hot spot is surrounded by a larger, warmer region. The large observed pulsed fractions are achievable in a realistically modeled NS atmosphere that accounts for the different opacities of the normal modes of polarization in a strong magnetic field (Özel, 2001; Özel et al., 2001). Many geometries were modeled by Özel (2001), but only total pulsed fractions were reported rather than detailed light curves. It is important to model the pulse profiles completely, since the number of peaks and their phase relationship depend in detail on the surface and viewing geometry. In addition, the relative effects of fan versus pencil beaming, as regulated by the anisotropic opacity in a strong magnetic field, allow the number of peaks in the light curve to differ from the number of hot spots on the surface. Whereas the pulsed fraction of the hot component in XTE J1810$``$197 is 53%, this is not high enough to require a single spot. It is possible that the flat interpulse region of the triangular, hard X-ray pulse is actually emission from an antipodal spot at large viewing angle. Özel et al. (2001) showed that for large magnetic fields, most of the flux would be in a broad fan beam, while Thompson et al. (2002) argue that cyclotron resonant scattering in the magnetosphere can significantly reduce this effect. Detailed physical modeling of the pulse profiles of XTE J1810$``$197 is needed before definite conclusions about the emission and viewing geometries can be drawn. ## 4. Conclusions and Future Work Four sets of XMM-Newton observations of XTE J1810$``$197 spanning a year, clarify several new behaviors that might apply to other AXPs and AXP-like objects: * a two-temperature blackbody model is arguably preferred over the standard blackbody+power law model, * the components of the compound spectrum are decaying at rates that differ significantly, and * the component temperatures of the blackbody emission are nearly constant, implying that the area of emission is steadily decreasing. Several outstanding questions remain that can be addressed by continuing X-ray observations of XTE J1810$``$197 over the next few years. Will the current spectrum evolve into the prior quiescent one? Is the cool quiescent spectrum still present so that it will begin to dominate over the warm temperature component? To what degree is the quiescent emission pulsed? Why do the temperatures hardly change with time, requiring the apparent decrease in area? What is the emission geometry on the NS and what can we learn from it? Future observations will test our predictions for the flux decay and pulsed fractions and allow a better understanding of the relationship between the spectral components and the pulse profile. Ultimately, the goal of this study is to disentangle the processes of thermal and non-thermal emission to which magnetars convert their energy, during outburst and quiescence. We thank Fred Jansen and Norbert Schartel for providing the three XMM-Newton Target of Opportunity observations of XTE J1810$``$197. XMM-Newton is an ESA science mission with instruments and contributions directly funded by ESA Member States and NASA. This research is supported by NASA grants NNG05GC43G.
warning/0506/hep-ph0506301.html
ar5iv
text
# The Renormalization-group Method Applied to Non-equilibrium Dynamics Teiji KunihiroYukawa Institute for Theoretical Physics, Kyoto University, Kyoto, 606-0952, Japan ## 1 Introduction The theory of the nonequilibrium dynamics may be regarded as a collection of the theory how to reduce the dynamics of many-body systems to ones with fewer variables . In fact, the Bogoliubov-Born-Green-Kirkwood-Yvon hierarchy can be reduced to the time-irreversible Boltzmann equation, which is given solely in terms of the single-particle distribution function for dilute gas systems. The derivation of Boltzmann equation by Bogoliubov shows that the dilute-gas dynamics as a dynamical system with many-degrees of freedom has an attractive manifold spanned by the one-particle distribution function, which is also an invariant manifold . Boltzmann equation in turn can be further reduced to the hydrodynamic equation (Euler or Navier-Stokes equation) by a perturbation theory like Chapman-Enskog method or Bogoliubov’s method. Langevin equation which may be time-irreversible can be reduced to the time-irreversible Fokker-Planck equation with a longer time scale than the scale in Langevin equation. Two basic ingredients are commonly seen in the reduction of dynamics, which are interrelated with but relatively independent of each other: (i) The reduced dynamics is characterized with a longer time scale than that appearing in the original (microscopic) evolution equation. (ii) The reduced dynamics is described by a time-irreversible equation even when the original microscopic equation is time-reversible. The fundamental problem in the theory of deriving kinetic or transport equations is to clarify the mechanism of and to implement the above two basic ingredients. In a few years ago, we showed that the so called the renormalization-group method, which is formulated in terms of the classical theory of envelopes, gives a unified theory for the reduction of the dynamics and can be used to derive the various transport equations in a transparent way; we have also elucidated that the underlying mathematics of the reduction is an explicit construction of the invariant manifold, a notion in dynamical system. In this report, we pick up the problem to derive the hydrodynamics equations from the Boltzmann equation as a typical problem of the reduction appearing in the field of non-equilibrium dynamics. Some remarks will be also given on applications of the method to other problems such as the reduction of the dynamics of Fokker-Planck equation and also on an extension to the relativistic case. ## 2 Derivation of hydrodynamic equations from Boltzmann equation In this section, we apply the RG method formulated in to derive Euler and Navier-Stokes equations, successively from the Boltzmann equation. ### 2.1 Basics of Boltzmann equation Boltzmann equation is an evolution equation of the one-particle distribution function $`f(𝒓,𝒗,t)`$ in the phase space, and reads $`{\displaystyle \frac{f}{t}}+𝒗{\displaystyle \frac{f}{𝒓}}=I[f].`$ (1) Here the left-hand side describes the change due to the canonical equation of motion while the right-hand side the change due to collisions; $`I[f]`$ $`=`$ $`{\displaystyle 𝑑𝒗_1𝑑𝒗^{}𝑑𝒗_1^{}w(𝒗𝒗_1|𝒗^{}𝒗_1^{})}`$ $`\times \{f(𝒓,𝒗^{},t)f(𝒓,𝒗_1^{},t)`$ (2) $``$ $`f(𝒓,𝒗,t)f(𝒓,𝒗_1,t)\},`$ which is called the collision integral. The transition probability $`w(𝒗𝒗_1|𝒗^{}𝒗_1^{})`$ has the following symmetry due to the time-reversal invariance of the microscopic equation of motion; $`w(𝒗𝒗_1|𝒗^{}𝒗_1^{})=w(𝒗^{}𝒗_1^{}|𝒗𝒗_1).`$ Furthermore, the invariance under the particle-interchange implies the following equality; $`w(𝒗𝒗_1|𝒗^{}𝒗_1^{})=w(𝒗_1𝒗|𝒗_1^{}𝒗^{})`$ $`=w(𝒗_1^{}𝒗^{}|𝒗_1𝒗)`$. We say that the function $`\phi (𝒗)`$ is a collision invariant if it satisfies the following equation; $`{\displaystyle 𝑑𝒗\phi (𝒗)I[f]}=0.`$ (3) Owing to the conservation of the particle number, the total momentum and the total kinetic energy during the collisions, the quantities given by linear combination of the following five quantities are collision invariant: $`1,𝒗`$ and $`v^2`$. For a collision invariant $`\phi (𝒗)`$, we can define the density $`n_\phi `$ and the current $`𝒋_\phi `$ as follows; $`n_\phi `$ $`=`$ $`{\displaystyle 𝑑𝒗\phi (𝒗)f(𝒓,𝒗,t)},`$ $`𝒋_\phi `$ $`=`$ $`{\displaystyle 𝑑𝒗𝒗\phi (𝒗)f(𝒓,𝒗,t)},`$ (4) which satisfy the continuity or balance equation; $`_tn_\phi +𝒋_\phi =0.`$ (5) Thus we have formally the hydrodynamic equations as the balance equations for the conservation of the particle number, total momentum and kinetic energy. These equations are, however, formal ones because the distribution function $`f`$ is not yet solved: The solution obtained from Boltzmann equation will give the explicit forms of the internal energy, the transport coefficients and so on. The $`H`$ function is defined as follows: $`H(𝒓,t)={\displaystyle 𝑑𝒗f(𝒓,𝒗,t)\left(\mathrm{ln}f(𝒓,𝒗,t)1\right)}.`$ (6) For equilibrium states, the $`H`$ function is equal to the entropy $`S`$ with the sign changed. Defining the corresponding current by $`𝑱_H(𝒓,t)={\displaystyle 𝑑𝒗𝒗f(𝒓,𝒗,t)\left(\mathrm{ln}f(𝒓,𝒗,t)1\right)},`$ (7) one has the balance equation; $`{\displaystyle \frac{H}{t}}+𝑱_H={\displaystyle 𝑑𝒗I[f]\mathrm{ln}f}.`$ (8) This shows that when $`\mathrm{ln}f`$ is a collision invariant, the $`H`$ function is conserved. In fact, one can show that this condition is satisfied when $`f(𝒓,𝒗,t)`$ is a local equilibrium distribution function as given by (19) below. ### 2.2 Application of the RG method to derive hydrodynamic equations To make it clear that the following discussion fits to the general formulation given in , we discretize the argument $`𝒗`$ as $`𝒗𝒗_i`$: Discriminating the arguments $`(𝒓,t)`$ and $`𝒗_i`$ in $`f(𝒓,𝒗_i,t)`$, we indicate $`𝒗_i`$ as a subscript $`i`$ for the distribution function; $`f(𝒓,𝒗_i,t)=f_i(𝒓,t)(𝒇(𝒓,t))_i.`$ Then Boltzmann equation now reads $`{\displaystyle \frac{f_i}{t}}=\widehat{I}[𝒇]_i𝒗_i{\displaystyle \frac{f_i}{𝒓}},`$ (9) where $`\widehat{I}[𝒇]_i={\displaystyle \underset{j,k,l}{}}w(𝒗_i𝒗_j|𝒗_k𝒗_l)(f_kf_lf_if_j)(𝒓,t).`$ (10) Now let us consider a situation where the fluid motion is slow with long wave-lengths so that $`𝒗_i{\displaystyle \frac{f_i}{𝒓}}=O(ϵ),`$ (11) where $`ϵ`$ is a small quantity,$`|ϵ|<1.`$ To take into account the smallness of $`ϵ`$ in the following calculations formally, let us introduce the scaled coordinate $`\overline{𝒓}`$ defined by $`\overline{𝒓}=ϵ𝒓`$ and $`/𝒓=ϵ/\overline{𝒓}`$. Then (9) reads $`{\displaystyle \frac{f_i}{t}}=\widehat{I}[𝒇]_iϵ𝒗_i{\displaystyle \frac{f_i}{\overline{𝒓}}},`$ (12) which has a form to which the perturbation theory given in is naturally applicable. In accordance with the general formulation given in , we first expand the solution as follows; $`f_i(\overline{𝒓},t)=f_i^{(0)}(\overline{𝒓},t)+ϵf_i^{(1)}(\overline{𝒓},t)+\mathrm{}.`$ Let $`\stackrel{~}{f}_i(\overline{𝒓},t;t_0))`$ be a solution around $`tt_0`$ given by a perturbation theory with $`f_i(\overline{𝒓},t_0)`$ being the initial value at $`t=t_0`$; $`\stackrel{~}{f}_i(\overline{𝒓},t=t_0;t_0)=f_i(\overline{𝒓},t_0).`$ (13) We expand $`\stackrel{~}{f}_i(\overline{𝒓},t;t_0)`$ as $`\stackrel{~}{f}_i(\overline{𝒓},t;t_0)`$ $`=`$ $`\stackrel{~}{f}_i^{(0)}(\overline{𝒓},t;t_0)+ϵ\stackrel{~}{f}_i^{(1)}(\overline{𝒓},t,t_0)`$ (14) $`+`$ $`\mathrm{},`$ and the respective initial condition is set up as follows, $`\stackrel{~}{f}_i^{(l)}(\overline{𝒓},t=t_0;t_0)=f_i^{(l)}(\overline{𝒓},t_0),(l=0,1,2\mathrm{}).`$ (15) The $`0`$-th order equation reads $`{\displaystyle \frac{\stackrel{~}{f}_i^{(0)}}{t}}=(I[\stackrel{~}{𝒇}^{(0)}])_i.`$ (16) Now we are interested in the slow motion which may be achieved asymptotically as $`t\mathrm{}`$. Therefore we take the following stationary solution, $`{\displaystyle \frac{\stackrel{~}{f}_i^{(0)}}{t}}=0,`$ (17) which is a fixed point of the equation satisfying $`(\widehat{I}[\stackrel{~}{𝒇}^{(0)}])_i=0,`$ (18) for arbitrary $`\overline{𝒓}`$. Notice that (18) shows that the distribution function $`\stackrel{~}{𝒇}^{(0)}`$ is a function of collision invariants. Such a distribution function is a local equilibrium distribution function or Maxwellian; $`\stackrel{~}{f}_i^{(0)}(\overline{𝒓},t;t_0)`$ $`=`$ $`n(\overline{𝒓},t_0)({\displaystyle \frac{m}{2\pi k_BT(\overline{𝒓},t_0)}})^{3/2}`$ (19) $`\times `$ $`\mathrm{exp}[{\displaystyle \frac{m|𝒗_i𝐮(\overline{𝒓},t_0)|^2}{2\pi k_BT(\overline{𝒓},t_0)}}].`$ Here, the local density $`n`$, local temperature $`T`$, local flux $`𝐮`$ are all dependent on the initial time $`t_0`$ and the space coordinate $`\overline{𝒓}`$ but independent of time $`t`$. The first-order equation reads $`(({\displaystyle \frac{}{t}}A)\stackrel{~}{f}^{(1)})_i=𝒗_i{\displaystyle \frac{\stackrel{~}{f}_i^{(0)}}{\overline{𝒓}}}.`$ (20) Here the linear operator $`A`$ is defined by $`[\widehat{I}^{}[\stackrel{~}{𝒇}^{(0)}]\stackrel{~}{𝒇}^{(1)}]_i`$ $`=`$ $`{\displaystyle \underset{j=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{I}{\stackrel{~}{f}_j}}|_{\stackrel{~}{𝒇}=\stackrel{~}{𝒇}^0}\stackrel{~}{f}_j^{(1)}`$ (21) $``$ $`(A\stackrel{~}{𝒇}^{(1)})_i.`$ Defining the inner product between $`𝝋`$ and $`𝝍`$ by $`𝝋,𝝍={\displaystyle 𝑑𝒗\phi \psi },`$ (22) one can show that $`A`$ is self-adjoint; $`\phi ,A\psi =A\phi ,\psi .`$ (23) One can further show that the five invariants $`m,𝒗,\frac{m}{2}v^2`$ span the kernel of $`A`$; $`\mathrm{KerA}=\{m,𝒗,\frac{m}{2}𝒗^2\}.`$ The other eigenvalues are found to be negative because one can show $`\phi ,A\phi 0.`$ (24) We write the projection operator to the kernel as P and define Q $`=1`$ P. Applying the general formulation in , one can readily obtain the first-order solution, $`\stackrel{~}{𝒇}^{(1)}=(tt_0)\mathrm{P}𝒗{\displaystyle \frac{\stackrel{~}{𝒇}^{(0)}}{\overline{𝒓}}}+A^1\mathrm{Q}𝒗{\displaystyle \frac{\stackrel{~}{𝒇}^{(0)}}{\overline{𝒓}}}.`$ (25) The perturbative solution up to the $`ϵ`$ order is found to be $`\stackrel{~}{𝒇}(\overline{𝒓},t,t_0)`$ $`=`$ $`\stackrel{~}{𝒇}^{(0)}(\overline{𝒓},t,t_0)+ϵ[(tt_0)\mathrm{P}𝒗{\displaystyle \frac{\stackrel{~}{𝒇}^{(0)}}{\overline{𝒓}}}`$ (26) $`+A^1\mathrm{Q}𝒗{\displaystyle \frac{\stackrel{~}{𝒇}^{(0)}}{\overline{𝒓}}}].`$ Notice the appearance of a secular term. If one stops the approximation and apply the RG equation, $`\stackrel{~}{𝒇}/t_0|_{t_0=t}=\mathrm{𝟎},`$ (27) one has $`{\displaystyle \frac{\stackrel{~}{𝒇}^{(0)}}{t}}+ϵ\mathrm{P}𝒗{\displaystyle \frac{\stackrel{~}{𝒇}^{(0)}}{\overline{𝒓}}}=\mathrm{𝟎}.`$ (28) This is a master equation from which equations governing the time evolution of $`n(𝒓,t)`$, $`𝐮(𝒓,t)`$ and $`T(𝒓,t)`$ in $`\stackrel{~}{𝒇}^{(0)}`$, i.e., fluid dynamical equations: In fact, taking an inner product between $`m`$, $`m𝒗`$ and $`mv^2/2`$ with this equation, one has fluid dynamical equations as given below. Notice that $`f`$ is now explicitly solved and the energy density $`e`$, the pressure tensor $`P_{ij}`$ and the heat flux $`Q_i`$ are given as follows; $`e(𝒓,t)`$ $`=`$ $`{\displaystyle 𝑑𝒗\frac{m}{2}|𝒗𝐮|^2f^{(0)}(𝒓,𝒗,t)}`$ (29) $`=`$ $`{\displaystyle \frac{3}{2}}k_BT(𝒓,t),`$ $`P_{ij}(𝒓,t)`$ $`=`$ $`{\displaystyle 𝑑𝒗m(v_iu_i)(v_ju_j)f^{(0)}(𝒓,𝒗,t)}`$ (30) $`=`$ $`nk_BT(𝒓,t)\delta _{ij}`$ $``$ $`P(𝒓,t)\delta _{ij},`$ $`Q_i(𝒓,t)`$ $`=`$ $`{\displaystyle 𝑑𝒗\frac{m}{2}|𝒗𝐮|^2(v_iu_i)f^{(0)}(𝒓,𝒗,t)}`$ (31) $`=`$ $`0.`$ We have defined the pressure $`P`$ using the equation of state for the ideal gas in the second line. There is no heat flux because the distribution function $`f^{(0)}`$ in the formulae is the one for the local equilibrium. Inserting these formulae into (5), we end up with a fluid dynamical equation without dissipation, i.e., the Euler equation; $`{\displaystyle \frac{\rho }{t}}`$ $`+`$ $`\rho 𝐮=0,`$ (32) $`{\displaystyle \frac{(\rho u_i)}{t}}`$ $`+`$ $`{\displaystyle \frac{}{x_j}}(\rho u_iu_j)+{\displaystyle \frac{}{x_i}}P=0,`$ (33) $`{\displaystyle \frac{}{t}}(\rho u^2+e)`$ $`+`$ $`{\displaystyle \frac{}{x_i}}[(\rho {\displaystyle \frac{u^2}{2}}+e+P)u_i]=0.`$ (34) We notice that these equations have been obtained from the RG equation (28). It should be emphasized however that the distribution function obtained in the present approximation takes the form $`f(𝒓,𝒗,t)`$ $`=`$ $`f^{(0)}(𝒓,𝒗,t)`$ (35) $`+`$ $`A^1\mathrm{Q}𝒗{\displaystyle \frac{f^{(0)}(𝒓,𝒗,t)}{𝒓}},`$ which incorporates as a perturbation a distortion from the local equilibrium distribution and gives rise to dissipations. One can proceed to the second order approximation straightforwardly and obtain fluid dynamical equation with dissipations as the RG equation. The perturbation equation in the second order reads $`(({\displaystyle \frac{}{t}}A)\stackrel{~}{f}^{(2)})_i=𝒗_i{\displaystyle \frac{\stackrel{~}{f}_i^{(1)}}{\overline{𝒓}}}.`$ (36) Here, we must make an important notice: We have actually used the linearized Boltzmann equation neglecting the second-order term of $`\stackrel{~}{𝒇}^{(1)}`$ in the collision integral: It is known that the neglected term produces the so called Burnett terms which are absent in the usual Navier-Stokes equation . Applying the RG method, we have $`{\displaystyle \frac{f^{(0)}}{t}}`$ $`+`$ $`ϵ\mathrm{P}𝒗f^{(0)}`$ (37) $`+ϵ^2\mathrm{P}𝒗A^1𝐐𝒗f^{(0)}=0.`$ The third term represents dissipations. Taking inner products between the first equation and the particle number, the velocity and the kinetic energy, one will obtain a fluid dynamical equation with dissipations included, i.e., Navier-Stokes equation. ## 3 Summary and concluding remarks In summary, we have shown that the fluid dynamical limit of the Boltzmann equation can be obtained neatly in the RG method as formulated in . It would be intriguing to see what kind of the dissipative fluid dynamical would emerge when the RG method is applied to the relativistic kinetic equations. The reduction of a kinetic equation to a further slower dynamics appear quite often, reflecting the hierarchy of the space-time of nature. A typical problem in this category is to derive Smoluchowski equation from Kramers equation. As is expected, it is possible to develop a systematic theory for the adiabatic elimination of fast variables in Fokker-Planck equations, which appear in the theory of the critical dynamics. We hope that we can report a development on these problems in near future.
warning/0506/astro-ph0506392.html
ar5iv
text
# Cosmological constraints on a classical limit of quantum gravity ## I Introduction The observed acceleration of the universe, confirmed from a cross-correlation of independent cosmological datasets Riess:1998cb ; Perlmutter:1998np ; Tonry:2003zg ; Bennett:2003bz ; Netterfield:2001yq ; Halverson:2001yy , poses a crucial question for fundamental physics. There are at least three distinct approaches from which an explanation of these observations might come. The simplest explanation is that cosmic acceleration is due to a cosmological constant. If this is the case, then our understanding of quantum fluctuations of matter fields and their gravitational effects will require new insights. A second possibility is that there is no cosmological term and that the dynamics of some new matter field, for example a scalar, lead to late-time acceleration. Such an explanation would require us to understand the existence of an unnaturally weakly coupled long-range field, with minute energy density at this particular epoch in cosmic history. A third, and less well-explored, possibility is that cosmic acceleration is due to an infrared modification of gravity, which yields cosmological solutions radically different from those of pure general relativity. While many such models exist Carroll:2003wy ; Capozziello:2003tk ; Carroll:2004de ; Deffayet:2001pu ; Freese:2002sq ; Arkani-Hamed:2002fu ; Dvali:2003rk ; Arkani-Hamed:2003uy ; Vollick ; Easson:2004fq ; Dick ; Nojiri\_1 ; Dolgov ; Nojiri\_2 ; Meng1 ; Chiba ; Meng2 ; Flanagan ; Woodard ; Ezawa ; Flanagan\_2 ; Rajaraman ; Vollick\_2 ; Nunez ; Allemandi ; Lue:2003ky , it has recently been shown by two of the authors that a unique theory is selected by the special geometry of Lorentzian spacetimes approximating a quantum spacetime Schuller:2004nn ; Schuller:2004rn . In this paper, we explore the cosmology of this distinguished deformation of Einstein gravity. The deformation takes the form of an inverse curvature term constructed from the Riemann tensor, similar to, but distinct from the modifications discussed in Carroll:2004de . In particular we investigate in some detail the phase space structure of the resulting modified Friedmann equation. In ordinary Einstein gravity, if the present acceleration of the Universe is attributed to a cosmological constant, then any cosmology inevitably approaches the de Sitter vacuum at late times. In our case, although the theory is constructed to have de Sitter space as a vacuum solution, and thus to be relevant to the accelerating universe, this solution is unstable, and the true phase space of cosmological solutions is significantly more complex, displaying two late-time power-law attractors – one accelerating and the other dramatically decelerating. However, it appears that the theory cannot explain the accelerating universe. This is because non-accelerating cosmologies sit on a separatrix between the basins of attraction of the two attractors. It is therefore impossible to pass from the decelerating cosmology required, in standard cosmology, by nucleosynthesis and structure formation to the accelerating one inferred from supernovae Ia. The supernovae data provides evidence for the jerk at greater than $`99\%`$ confidence level Carroll:2001bv . The phase space is divided into ever accelerating and ever decelerating universes; cross-overs cannot occur. However, for alternative cosmologies, such as the one under investigation here, the crossover requirement is less clear and there might be some loopholes which we will discuss. The structure of this paper is as follows. In the next section we describe inverse Riemann gravity and then derive the cosmological equations of motion necessary for the remainder of the paper. In section III we provide an analytic phase space description of the most important features of the solutions. We show that the de Sitter solution is unstable, identify late-time power-law attractors and, most importantly, demonstrate the existence of a critical zero-acceleration separatrix. In section IV we go on to provide complete numerical solutions to the equations of motion, mapping out the phase space for both the vacuum and non-vacuum cases, before concluding with a discussion in section V. ## II Inverse Riemann cosmology ### II.1 Inverse Riemann gravity Quantum gravity heuristics suggest that Lorentzian manifolds approximating a quantum spacetime possess both infrared and ultraviolet sectional curvature bounds Schuller:2004nn ; Schuller:2004rn . Such bounds are entirely equivalent to bounds on the eigenvalues of the Riemann tensor, considered as an endomorphism on the space of 2-forms. If one attempts to write down an action such that all solutions respect these bounds, one quickly discovers that there is a large class of such theories (as many as there are holomorphic functions on an annulus). The ambiguity arises because of our ignorance of the exact quantum spacetime structure in this approach. However, because we are interested in cosmology, it is appropriate to take the classical limit, corresponding to the removal of the curvature bounds. If one further requires that, in this limit, de Sitter spacetime is a solution, as one might hope occurs in order to explain the accelerating universe, then the resulting classical theory is unique <sup>5</sup><sup>5</sup>5To be more precise, uniqueness of the classical limit holds within a (more tractable) subclass of all possible theories respecting the sectional curvature bounds, for which the Laurent series in the Lagrangian is already determined by a set of Taylor coefficients Schuller:2004nn .. Surprisingly, and interestingly, it is not the Einstein-Hilbert action with a positive cosmological constant. In this paper, we take the resulting classical action as our starting point, and proceed to investigate its broad cosmological characteristics. The action is a new type of infrared modification of Einstein-Hilbert gravity. Including a matter action $`S_m`$, the total action in $`d`$ spacetime dimensions reads $$S=d^dx\sqrt{g}\text{ Tr}\left(R\frac{d2}{d+2}\zeta ^2R^1\right)+S_m,$$ (1) where $`R^A_B`$ is the endomorphism on the space of antisymmetric two-tensors defined by the Riemann tensor $`R^{[ab]}_{[cd]}`$. The trace should not double-count, and hence, is defined by $`\mathrm{Tr}f(R)f(R)_{}^{A}{}_{A}{}^{}=f(R)_{}^{[ab]}{}_{[ab]}{}^{}/2`$. We have chosen units such that the reduced Planck mass is $`M_P=(8\pi G)^{1/2}=1`$. The vacuum solutions are de Sitter and anti de Sitter space of constant curvature, proportional to $`\pm \zeta `$ respectively. The action (1) is similar to that introduced in Carroll:2003wy . However, note that the current action contains both the Ricci scalar and a specific function of the full Riemann tensor, rather than just a function of the Ricci scalar. In particular, it is important to note that $`\mathrm{Tr}(R^1)`$ is generally not equal to the inverse of the Ricci scalar. As noted in Schuller:2004nn , the dimensionality of the space of metrics is different from that of the space of Riemann tensors, which means that it is not possible to perform a transformation of the metric, such as a conformal transformation, which reduces the action to that of Einstein gravity, with the extra degrees of freedom represented by additional matter fields. This means that direct comparison with solar system tests is a more complicated calculational task, which we will not attempt in this paper. It thus remains to be seen whether this action is compatible with gravitational dynamics at the solar system and galactic scales. The general equation of motion resulting from (1) is $$R^{ij}\frac{1}{2}g^{ij}R+\frac{d2}{d+2}\zeta ^2\left[(R^1)^{(i|b|j)}{}_{b}{}^{}+\frac{1}{2}g^{ij}(R^1)^{ab}{}_{ab}{}^{}_b_c(R^2)^{c(ij)b}\right]=T^{ij},$$ (2) where $`T^{ij}`$ is the matter energy-momentum tensor derived from $`S_m`$, and the sign convention $`R^a{}_{bcd}{}^{}=_c\mathrm{\Gamma }^a{}_{bd}{}^{}+\mathrm{\Gamma }^a{}_{ec}{}^{}\mathrm{\Gamma }_{}^{e}{}_{bd}{}^{}(cd)`$ has been used. ### II.2 Cosmological equations of motion We impose the cosmological Friedmann-Lemaître-Robertson-Walker ansatz $$ds^2=dt^2+a(t)^2d\mathrm{\Sigma }_k^2$$ (3) for a homogenous and isotropic spacetime, where $`a(t)`$ is the scale factor and $$d\mathrm{\Sigma }_k^2=\overline{g}_{\alpha \beta }dx^\alpha dx^\beta =\frac{dr^2}{1kr^2}+r^2\left(d\theta ^2+\mathrm{sin}^2\theta d\varphi ^2\right)$$ (4) is the metric on a $`d1`$ dimensional space of normalized constant curvature $`k=0,\pm 1`$ with Riemann tensor $`\overline{R}_{\alpha \beta \gamma \delta }=k(\overline{g}_{\alpha \gamma }\overline{g}_{\beta \delta }\overline{g}_{\alpha \delta }\overline{g}_{\beta \gamma })`$. With this ansatz, $`R^A_B`$ is a diagonal matrix, with indices taking values in $`\{[0\beta ],[\alpha \beta ]\}`$. The corresponding $`d(d1)/2`$ eigenvalues are given by $`R^{[0\beta ]}_{[0\delta ]}`$ $`=`$ $`{\displaystyle \frac{\ddot{a}}{a}}\delta _\delta ^\beta ,`$ (5) $`R^{[\alpha \beta ]}_{[\gamma \delta ]}`$ $`=`$ $`{\displaystyle \frac{(k+\dot{a}^2)}{a^2}}\left(\delta _\gamma ^\alpha \delta _\delta ^\beta \delta _\delta ^\alpha \delta _\gamma ^\beta \right),`$ (6) the overdot denoting differentiation with respect to cosmic time $`t`$. Thus, the inverse curvature correction to the Einstein-Hilbert action appearing in (1) is given by $$\mathrm{Tr}(R^1)=(d1)\left[\left(\frac{d2}{2}\right)\frac{a^2}{k+\dot{a}^2}+\frac{a}{\ddot{a}}\right].$$ (7) From the form of this correction we may already anticipate a possible difficulty in passing from a decelerating phase to an accelerating one, since (7) typically blows up when $`\ddot{a}=0`$. We consider perfect fluid matter with the energy-momentum tensor $$T^{ij}=(\rho +p)u^iu^j+pg^{ij},$$ (8) where $`u^i`$ characterizes the fluid’s four-velocity in its rest frame, $`\rho `$ denotes its energy density, and $`p`$ its pressure. The analogue of the Friedmann equation is derived from the time-time component of the equations of motion which, after some algebra, becomes $`{\displaystyle \frac{2\zeta ^2}{d+2}}[{\displaystyle \frac{2a}{\ddot{a}}}+d\left({\displaystyle \frac{\dot{a}}{\ddot{a}}}\right)^2{\displaystyle \frac{2a\dot{a}\stackrel{\dot{}\dot{}\dot{}}{a}}{\ddot{a}^3}}`$ $``$ $`{\displaystyle \frac{(d2)a^2(k+3\dot{a}^2)}{2(k+\dot{a}^2)^2}}]`$ (9) $`+`$ $`{\displaystyle \frac{k+\dot{a}^2}{a^2}}={\displaystyle \frac{2\rho }{(d1)(d2)}}.`$ Note that this result, from the full equations of motion, agrees with an effective action calculation performed with the line element (3) additionally including a lapse function $`N(t)`$ which is set to unity after variation. Setting $`\zeta 0`$, one obtains the usual Friedmann equation of the standard cosmology based on the Einstein-Hilbert action. For later convenience, we adopt the usual definition of the Hubble parameter $`H(t)\dot{a}(t)/a(t)`$, which allows us to rewrite the above equation in the form $`{\displaystyle \frac{2\zeta ^2}{d+2}}[{\displaystyle \frac{2}{\dot{H}+H^2}}+d\left({\displaystyle \frac{H}{\dot{H}+H^2}}\right)^2`$ $``$ $`{\displaystyle \frac{2H(\ddot{H}+3\dot{H}H+H^3)}{(\dot{H}+H^2)^3}}{\displaystyle \frac{(d2)a^2(k+3a^2H^2)}{2(k+a^2H^2)^2}}]`$ (10) $`+`$ $`{\displaystyle \frac{k+a^2H^2}{a^2}}={\displaystyle \frac{2\rho }{(d1)(d2)}}.`$ This equation and the continuity equation $$\dot{\rho }+(d1)H(\rho +p)=0,$$ (11) which follows directly from the conservation of energy-momentum $`_aT^{ab}=0`$, are equivalent to the full set of gravitational field equations. ## III Phase Space Description of Solutions The main goal of this work is to understand the expansion history of the universe in the theory introduced in the previous section. Because the resulting system of equations is of order greater than two, it is most convenient to understand the generic properties by considering its phase space evolution. As is typical in cosmology, we close the system of equations by specifying a matter equation of state $`p=p(\rho )`$. Then, without further assumptions, the cosmological phase space of the resulting system is given by $`(\rho ,a,\dot{a},\ddot{a})`$. The diffeomorphism invariance of the field equations (in particular, the time reparametrization invariance) bundles the trajectories in this phase space into classes of physically indistinguishable solutions. We may pick one member of each class by requiring that for a fixed time $`t_0`$ the scale factor should satisfy $`a(t_0)=1`$. Note that this does not change the dimension of the phase space; it merely implies that, in finding solutions, one does not need to scan over different initial conditions for the scale factor $`a`$. Note that upon a global conformal rescaling $`g_{ij}\xi ^1g_{ij}`$, the Riemann endomorphism scales with $`\xi `$ and the metric density with $`\xi ^{d/2}`$. This allows to factor out $`\xi ^{(2d)/2}`$ while simultaneously rescaling the deformation parameter $`\zeta \zeta /\xi `$. For a matter action that also scales with $`\xi ^{(2d)/2}`$, such as a massless scalar or simply the vacuum, solutions of theories with different non-zero $`\zeta `$ are therefore related by global conformal transformations. This means that the qualitative nature of the phase space will be the same for all values of $`\zeta `$. We now survey the salient features of the phase space. ### III.1 De Sitter vacuum instability The equations of motion (2) have precisely two solutions of constant curvature, which are the de Sitter and anti-de Sitter spacetimes with Hubble parameter (in the flat slicing) $`\pm \zeta `$. Note that this statement is independent of the spacetime dimension $`d`$. For convenience we define $`J\dot{H}`$ and solve equation (10) with $`k=0`$ for $`\ddot{H}`$. This yields the system of equations $`\dot{\rho }`$ $`=`$ $`(d1)H\rho ,`$ (12) $`\dot{H}`$ $`=`$ $`J,`$ (13) $`\dot{J}`$ $`=`$ $`{\displaystyle \frac{(J+H^2)^3}{2H}}\{{\displaystyle \frac{d+2}{2\zeta ^2}}[H^2{\displaystyle \frac{2\rho }{(d1)(d2)}}]{\displaystyle \frac{2}{J+H^2}}`$ (14) $`+{\displaystyle \frac{dH^2}{(J+H^2)^2}}{\displaystyle \frac{2H(3HJ+H^3)}{(J+H^2)^3}}{\displaystyle \frac{3(d2)}{2H^2}}\}.`$ This system possesses two fixed points. The first is at $`(\rho ,H,J)=(0,\sqrt{\zeta },0)`$, and coincides precisely with the de Sitter vacuum solution $`a(t)=\mathrm{exp}\sqrt{\zeta }t`$. The second fixed point is $`(\rho ,H,J)=(\rho _0,0,0)`$ for constant $`\rho _0`$. However, this fixed point satisfies $`J+H^2=0`$, which means that it occurs at a singularity of equation (10). Therefore this is not an admissible solution. To analyze the stability properties of the de Sitter fixed point under cosmological evolution, we linearize the above equations setting $`(\rho ,H,J)=(x,\sqrt{\zeta }+y,z)`$ with small $`X^T=(x,y,z)`$. As we shall see, the de Sitter solution is a saddle point in phase space, and hence unstable. The linearized system can be written in the form $`\dot{X}=MX`$ with the matrix $`M`$ given by $$M=\left(\begin{array}{ccc}(d1)\sqrt{\zeta }& 0& 0\\ 0& 0& 1\\ \frac{(d+2)\sqrt{\zeta }}{2(d1)(d2)}& (d+2)\zeta & (d1)\sqrt{\zeta }\end{array}\right).$$ (15) The eigenvalues of this matrix determine the type of fixed point; they are $`\lambda _1`$ $`=`$ $`(d1)\sqrt{\zeta },`$ (16) $`\lambda _\pm `$ $`=`$ $`{\displaystyle \frac{1}{2}}\sqrt{\zeta }\left(d+1\pm \sqrt{d^2+2d+9}\right).`$ (17) Clearly, $`\lambda _1<0`$, implying the stability of the de Sitter solution against perturbations of the energy density $`\rho `$ away from the vacuum value. However, $`\lambda _+>0`$ and $`\lambda _{}<0`$, implying that this solution is a saddle point, having one unstable direction in phase space. This behaviour can be seen clearly in the numerical phase space plots presented below. ### III.2 Power law attractors The simplest form of accelerating expansion in the final stage of the cosmological evolution, namely a stable de Sitter vacuum, is ruled out, as we have seen. However, there exist power law attractors. Substituting the ansatz $`at^\gamma `$, with constant $`\gamma `$, into the modified Friedmann equation (9), one finds late-time solutions ($`t\mathrm{}`$). For $`k=0`$, in $`d>2`$, we have $$\gamma _\pm =\frac{1}{d+2}\left(3d\pm \sqrt{6(d^2+2)}\right).$$ (18) In $`d=4`$, these exponents are $$\gamma _\pm ^{(d=4)}=2\pm \sqrt{3},$$ (19) and correspond to effective equation of state parameters $$w_{\mathrm{eff}}^\pm =\frac{12\sqrt{3}}{3}.$$ (20) We see that $`\gamma _+>1`$, which describes an accelerating solution, and $`0<\gamma _{}<1`$, which describes a decelerating solution. For $`k0`$, there is only the accelerating late-time solution $`\gamma _+`$, since the curvature term ultimately dominates, although, if inflation vastly flattens space at early times, then the decelerating behavior during the rest of cosmic history may include an intermediate regime approximated by the decelerating solution before curvature domination in the far future. A simple linearized analysis about these solutions shows that they are both attractors, with perturbations from them decaying away at late times. The saddle point instability of de Sitter is directly connected to the existence of the accelerating power law attractor. As we will see, a large class of cosmological phase space trajectories will come close to the de Sitter vacuum for a time, but ultimately approach power law behavior. ### III.3 The acceleration separatrix Perhaps the most cosmologically relevant feature of the modified Friedmann equation (9) is the generic singularity at $`\ddot{a}=0`$. This is, in fact, the location of a phase space separatrix. To see this, note that near $`\ddot{a}=0`$, equation (9) takes the asymptotic form $$\frac{d}{dt}\ddot{a}(t)2H(t)\ddot{a}(t).$$ (21) Since the Hubble parameter is always positive, it is clear that already accelerating solutions (with $`\ddot{a}`$ small and positive) are driven to accelerate even more, and already decelerating ones (with $`\ddot{a}`$ small and negative) are driven to further deceleration. Thus, phase-space trajectories can never cross over from deceleration to acceleration. This seems to create an almost insurmountable obstacle to the construction of a realistic cosmology in which the late-time acceleration of the universe is explained by the modified gravity nature of this theory. In standard cosmology, successful nucleosynthesis and structure formation require a decelerating cosmology at early times and hence a transition from deceleration to acceleration is required by observations. It is worth mentioning, however, that much of the intuition we have from standard cosmology is based on the implicit assumption that the Einstein equations (via the unmodified Friedmann equation) control the background dynamics. If instead, a modified Friedmann equation is used, many calculations beyond the scope of this paper must be redone in order to conclusively rule out a modified gravity theory. These include structure formation simulations and the determination of cosmological parameters, such as the total matter density, from a fit of the modified model to the cosmic microwave background. One clear example in which the standard wisdom proves inapplicable is the following: successful nucleosynthesis is possible even in an accelerating background if there is an asymmetry in the neutrino and antineutrino abundances. This is shown in Carroll:2001bv , without assuming a specific background dynamics. ## IV Numerical solutions While the asymptotic behavior of the theory can easily be analyzed analytically, we must solve the full equations numerically. To get a feeling for how solutions look, we restrict our attention to the case of $`d=4`$, flat $`k=0`$ cosmologies which have the most interesting phase space features. We first plot solutions to the vacuum equations of motion, setting $`\rho =p=0`$. Figure 1 shows solutions in the $`(\dot{H},H)`$ phase space. In the phase space plot we have scaled $`\zeta `$ so that the unstable de Sitter solution is located at the point $`(0,1)`$. The two late-time power law attractors and the separatrix corresponding to $`\ddot{a}=0`$ are clearly visible in the upper left quadrant of the phase space. Solutions with power-law behavior ($`at^\gamma `$), follow the curves $`H=\sqrt{\gamma \dot{H}}`$ in the phase space. The upper right quadrant of the phase space is included for completeness, however, these solutions have $`\dot{H}>0`$ and are physically irrelevant for the purposes of our paper. Since we are interested in the late universe, we now specialize to the case in which the dominant matter component is pure dust, for which the pressure vanishes $`p=0`$. We set $`\mathrm{\Omega }_m^{(0)}=0.28`$, $`a_0=1`$ and $`H(t=t_0)=H_0`$, where a subscript $`0`$ denotes the value of a quantity today. Figure 2 shows these solutions in the same $`(\dot{H},H)`$ phase space. We focus on solutions with initial conditions $`\rho (t_0)`$ and $`\dot{a}(t_0)`$, obtained from observations of the current values of the Hubble parameter and the density parameter $`\mathrm{\Omega }_mH^2\rho /(d1)`$, consistent with the present state of the universe within the context of Einstein gravity. We therefore set $`\rho (t_0)=3H_0^2\mathrm{\Omega }_m,a(t_0)=1`$ and $`\dot{a}(t_0)=H_0`$. It is then sufficient to scan over initial conditions for the remaining phase space variable $`\ddot{a}`$. Since we are in four dimensions, we only exhibit projections to three-dimensional hypersurfaces of constant $`a1`$. For the present values of the matter density and the Hubble parameter we choose $`\mathrm{\Omega }_m=0.28`$ and $`H_0=1`$. To find initial conditions for $`\ddot{a}`$ (or equivalently $`\dot{H}`$) compatible with the recent supernovae data, we assume Einstein gravity and a flat universe filled with dust ($`\mathrm{\Omega }_m`$) and a “dark energy” component with effective instantaneous equation of state parameter $`w_0`$. From the Einstein equations we find $$\dot{H}=\frac{3}{2}H^2\left(1+\frac{w_0}{1+\alpha }\right),$$ (22) where $$\alpha =\frac{\mathrm{\Omega }_m}{1\mathrm{\Omega }_m}.$$ (23) The initial conditions for $`\dot{H}`$, compatible with supernovae data are located in the green line in Figure 2, and are obtained using (22) together with the observational data $`1.48<w_0<0.72`$ at the $`95\%`$ confidence level astro-ph/0205096 ; astro-ph/0211522 ; Tonry:2003zg ; astro-ph/0407259 . ## V Conclusions The discovery of cosmic acceleration seems to imply that there is something important that we do not understand about particle physics, gravity or the interaction between them. While there have been many attempts to understand this phenomenon at a phenomenological level, it would be extremely exciting to derive the late-time evolution of the universe from a fundamental theory. In this paper we have investigated the cosmological evolution of the classical modified gravity theory given by the action (1). This theory arises in the following way: considerations of quantum gravity heuristics suggest that Lorentzian manifolds approximating a quantum spacetime possess sectional curvature bounds Schuller:2004nn ; Schuller:2004rn . A large class of dynamics whose solutions respect these curvature bounds exists. For a certain subclass, the classical limit corresponding to the removal of the curvature bounds, leads to an interesting result: one obtains a unique theory containing de Sitter space as a vacuum solution. The action for this theory is not the Einstein-Hilbert action with a positive cosmological constant, but rather contains a correction in terms of the inverse eigenvalues of the Riemann tensor. Using the resulting action as our starting point, we have investigated its broad cosmological characteristics. As is typical of modified gravity theories with inverse curvature terms Carroll:2004de ; Easson:2004fq , the equations of motion contain terms that are singular for certain values of the metric components. In the cosmological setting this translates into certain combinations of time derivatives of the scale factor. In many models Carroll:2004de ; Easson:2004fq these singularities lie in regions of the parameter space that are never reached by trajectories corresponding to realistic cosmological evolution. However, in the specific theory defined by the action (1), the relevant singularity of the equations of motion occurs precisely when $`\ddot{a}(t)`$=0. If one interprets the available cosmological data sets under the assumption that Einstein gravity controls the dynamics of the background, there is good evidence that the universe underwent a decelerating phase (needed for successful nucleosynthesis and structure formation), and now a late-time accelerating phase (to fit, among other data, the observations of type Ia supernovae). Thus standard cosmology requires that the universe pass from deceleration to acceleration at some point in cosmic history. This is not allowed by the singularity structure of the theory investigated here. Assuming that, even with the modified Friedmann equation (9), nucleosynthesis and structure formation can only take place during a decelerating phase, as in the standard Big-Bang model, the exclusion of cross-overs from decelerated to accelerated expansion presents an insurmountable obstacle to a viable cosmology. It is worth commenting, however, that in order to reliably confront alternative cosmologies with observations, one must either use data without assuming specific background dynamics, as is done e.g. in Carroll:2001bv ; Visser:2004bf , or redo standard calculations (such as extracting the cosmological parameters from data on the basis of the modified dynamics). This of course presents a formidable task and is beyond the scope of this paper. ###### Acknowledgements. The authors thank Joshua Goldberg, Stefan Hofmann and Wilfried Buchmüller for helpful comments. The work of MT and DE is supported in part by the National Science Foundation under grant PHY-0354990, by funds provided by Syracuse University and by Research Corporation.
warning/0506/math0506186.html
ar5iv
text
# 1 Non-Colliding Brownian Motions ## 1 Non-Colliding Brownian Motions By virtue of the Karlin-McGregor formula , the transition density of the absorbing Brownian motion in a Weyl chamber $$_<^N=\{𝐱=(x_1,x_2,\mathrm{},x_N)^N;x_1<x_2<\mathrm{}<x_N\},$$ is given by $$f_N(t;𝐱,𝐲)=\underset{1i,jN}{det}\left[p_t(x_i,y_j)\right],𝐱,𝐲_<^N,t[0,\mathrm{}),$$ where $`p_t`$ is the heat-kernel given by $$p_t(x,y)=\frac{1}{\sqrt{2\pi t}}e^{(xy)^2/2t}.$$ The integral $$𝒩_N(t;𝐱)=_{_<^N}f_N(t;𝐱,𝐲)𝑑𝐲,$$ where $`d𝐲=_{i=1}^Ndy_i`$, gives the probability that a Brownian motion starting from $`𝐱_<^N`$ does not hit the boundary of $`_<^N`$ up to time $`t>0`$. For a given $`T>0`$, we define (1.1) $$g_{N,T}(s,𝐱;t,𝐲)=\frac{f_N(ts;𝐱,𝐲)𝒩_N(Tt;𝐲)}{𝒩_N(Ts;𝐱)}$$ for $`0stT,𝐱,𝐲_<^N`$. It can be regarded as the transition probability density from the state $`𝐱_<^N`$ at time $`s`$ to the state $`𝐲_<^N`$ at time $`t`$. In it was shown that as $`|𝐱|0`$, $`g_N^T(0,𝐱;t,𝐲)`$ converges to (1.2) $$g_{N,T}(0,\mathrm{𝟎};t,𝐲)C(N,T,t)h_N(𝐲)\underset{i=1}{\overset{N}{}}p_t(0,y_i)𝒩_N(Tt,𝐲),$$ where $`C(N,T,t)=\pi ^{N/2}\left(_{j=1}^N\mathrm{\Gamma }(j/2)\right)^1T^{N(N1)/4}t^{N(N1)/2}`$, and $$h_N(𝐱)=\underset{1i<jN}{}(x_jx_i)=\underset{1i,jN}{det}\left[x_j^{i1}\right].$$ The $`N`$ particle system of non-colliding Brownian motions $`𝐗(t)`$ all started from the origin at time $`0`$, i.e. $`𝐗(0)=\mathrm{𝟎}=(0,0,\mathrm{},0)`$, and conditioned not to collide with each other in a time interval $`(0,T]`$ is defined by the process associated with the transition probability density $`g_{N,T}`$ given by (1.1) and (1.2). This process is temporally inhomogeneous and it was obtained as a diffusion scaling limit of the vicious walker model in . Figure 1 illustrates the process $`𝐗(t)`$. Let $`𝔛`$ be the space of countable subset $`\xi `$ of $``$ satisfying $`\mathrm{}(\xi K)<\mathrm{}`$ for any compact subset $`K`$. For $`𝐱=(x_1,x_2,\mathrm{},x_n)_{\mathrm{}=1}^{\mathrm{}}^{\mathrm{}}`$, we denote $`\{x_i\}_{i=1}^n𝔛`$ simply by $`\{𝐱\}`$. The diffusion process $`\{𝐗(t)\},t[0,T]`$ on the set $`𝔛`$ is defined by the transition probability density $`𝔤_{N,T}(s,\{𝐱\};t,\{𝐲\})`$, $`0stT`$: $$𝔤_{N,T}(s,\{𝐱\};t,\{𝐲\})=\{\begin{array}{cc}g_{N,T}(s,𝐱;t,𝐲),\hfill & \text{if}s>0,\mathrm{}\{𝐱\}=\mathrm{}\{𝐲\}=N,\hfill \\ g_{N,T}(0,\mathrm{𝟎};t,𝐲),\hfill & \text{if}s=0,\{𝐱\}=\{0\},\mathrm{}\{𝐲\}=N,\hfill \\ 0,\hfill & \text{otherwise},\hfill \end{array}$$ where $`𝐱`$ and $`𝐲`$ are the elements of $`_<^N`$. For the given time interval $`[0,T]`$, we consider the $`M`$ intermediate times $`t_0=0<t_1<t_2<\mathrm{}<t_M<t_{M+1}=T`$. The multi-time transition density of the process $`\{𝐗(t)\}`$ is given by $$𝔤_{N,T}(t_0,\{𝐱^{(0)}\};t_1,\{𝐱^{(1)}\};\mathrm{};t_{M+1},\{𝐱^{(M+1)}\})=\underset{\mu =0}{\overset{M}{}}𝔤_{N,T}(t_\mu ,\{𝐱^{(\mu )}\};t_{\mu +1},\{𝐱^{(\mu +1)}\}).$$ Here we set $`t_0=0`$, $`t_{M+1}=T`$ and $`\{𝐱^{(0)}\}\{0\}`$. From (1.1) and (1.2) we have (1.3) $`𝔤_{N,T}(0,\{0\};t_1,\{𝐱^{(1)}\};\mathrm{};t_{M+1},\{𝐱^{(M+1)}\})`$ $`=C(N,T,t_1)h_N\left(𝐱^{(1)}\right)\mathrm{sgn}\left(h_N\left(𝐱^{(M+1)}\right)\right)`$ $`\times {\displaystyle \underset{i=1}{\overset{N}{}}}p_{t_1}(0,x_i^{(1)}){\displaystyle \underset{\mu =1}{\overset{M}{}}}\underset{1i,jN}{det}\left[p_{t_{\mu +1}t_\mu }(x_i^{(\mu )},x_j^{(\mu +1)})\right].`$ For $`𝐱^{(\mu )}_<^N`$, $`1\mu M+1`$, and $`N^{}=1,2,\mathrm{},N`$, we write $`𝐱_N^{}^{(\mu )}=(x_1^{(\mu )},x_2^{(\mu )},\mathrm{},x_N^{}^{(\mu )})`$. For a sequence $`\{N_\mu \}_{\mu =1}^{M+1}`$ of positive integers less than or equal to $`N`$, we define the multi-time correlation function by (1.4) $`\rho _{N,T}(t_1,\{𝐱_{N_1}^{(1)}\};t_2,\{𝐱_{N_2}^{(2)}\};\mathrm{};t_{M+1},\{𝐱_{N_{M+1}}^{(M+1)}\})`$ $`={\displaystyle \underset{_{\mu =1}^{M+1}^{NN_\mu }}{}}{\displaystyle \underset{\mu =1}{\overset{M+1}{}}}{\displaystyle \frac{1}{(NN_\mu )!}}{\displaystyle \underset{i=N_\mu +1}{\overset{N}{}}}dx_i^{(\mu )}`$ $`𝔤_{N,T}(0,\{0\};t_1,\{𝐱_N^{(1)}\};t_2,\{𝐱_N^{(2)}\};\mathrm{};t_{M+1},\{𝐱_N^{(M+1)}\}).`$ Let $`C_0()`$ be the set of all continuous real functions with compact supports. For $`𝐟=(f_1,f_2,\mathrm{},f_{M+1})C_0()^{M+1}`$, and $`𝜽=(\theta _1,\theta _2,\mathrm{},\theta _{M+1})^{M+1}`$, the multi-time characteristic function is defined for the process $`𝐗(t)=(X_1(t),X_2(t),\mathrm{},X_N(t))`$ as $$\mathrm{\Psi }_{N,T}(𝐟;𝜽)=𝐄_{N,T}\left[\mathrm{exp}\left\{\sqrt{1}\underset{\mu =1}{\overset{M+1}{}}\theta _\mu \underset{i_\mu =1}{\overset{N}{}}f_\mu (X_{i_\mu }(t_\mu ))\right\}\right],$$ where $`𝐄_{N,T}[]`$ denotes the expectation determined by $`𝔤_{N,T}`$. Let $$\chi _\mu (x)=e^{\sqrt{1}\theta _\mu f_\mu (x)}1,1\mu M+1,$$ then by the definition of multi-time correlation function (1.4), we have (1.5) $`\mathrm{\Psi }_{N,T}(𝐟;𝜽)`$ $`=`$ $`{\displaystyle \underset{N_1=0}{\overset{N}{}}}{\displaystyle \underset{N_2=0}{\overset{N}{}}}\mathrm{}{\displaystyle \underset{N_{M+1}=0}{\overset{N}{}}}{\displaystyle _{^{N_1}}}𝑑𝐱_{N_1}^{(1)}{\displaystyle _{^{N_2}}}𝑑𝐱_{N_2}^{(2)}\mathrm{}{\displaystyle _{^{N_{M+1}}}}𝑑𝐱_{N_{M+1}}^{(M+1)}`$ $`{\displaystyle \underset{\mu =1}{\overset{M+1}{}}}{\displaystyle \underset{i_\mu =1}{\overset{N_\mu }{}}}\chi _\mu (x_{i_\mu }^{(\mu )})\rho _{N,T}(t_1,\{𝐱_{N_1}^{(1)}\};t_2,\{𝐱_{N_2}^{(2)}\};\mathrm{};t_{M+1},\{𝐱_{N_{M+1}}^{(M+1)}\}).`$ ## 2 Fredholm Pfaffian Representation of Characteristic Function Let $`Z_{N,T}[\chi ]`$ $`=`$ $`\left({\displaystyle \frac{1}{N!}}\right)^{M+1}{\displaystyle _{^{N(M+1)}}}{\displaystyle \underset{\mu =1}{\overset{M+1}{}}}d𝐱^{(\mu )}\underset{1i,jN}{det}\left[M_{i1}(x_j^{(1)})p_{t_1}(0,x_j^{(1)})(1+\chi _1(x_j^{(1)}))\right]`$ $`\times {\displaystyle \underset{\mu =1}{\overset{M}{}}}\underset{1i,jN}{det}\left[p_{t_{\mu +1}t_\mu }(x_i^{(\mu )},x_j^{(\mu +1)})(1+\chi _{\mu +1}(x_j^{(\mu +1)}))\right]\mathrm{sgn}\left(h_N(𝐱_N^{(M+1)})\right).`$ Here $`M_i(x)`$ is an arbitrary polynomial of $`x`$ with degree $`i`$ in the form $`M_i(x)=b_ix^i+\mathrm{}`$ with a constant $`b_i`$ for $`i=0,1,2,\mathrm{}`$, and thus $`h_N(𝐱)=det_{1i,jN}\left[M_{i1}(x_j)\right]/_{k=1}^Nb_{k1}`$. Then (1.3) gives (2.1) $$\mathrm{\Psi }_{N,T}(𝐟;𝜽)=\frac{Z_{N,T}[\chi ]}{Z_{N,T}[0]},$$ where $`Z_{N,T}[0]`$ is obtained from $`Z_{N,T}[\chi ]`$ by setting $`\chi _\mu (x)=0`$ for all $`\mu `$. By the well-known formula $$_{_<^N}𝑑𝐱\underset{1i,jN}{det}\left[\varphi _i(x_j)\right]\underset{1i,jN}{det}\left[\overline{\varphi }_i(x_j)\right]=\underset{1i,jN}{det}\left[_{}𝑑x\varphi _i(x)\overline{\varphi }_j(x)\right]$$ for square integrable continuous functions $`\varphi _i,\overline{\varphi }_i,1iN`$, we have $`Z_{N,T}[\chi ]`$ $`=`$ $`{\displaystyle _{_<^N}}d𝐲\underset{1i,jN}{det}[{\displaystyle _{^{M+1}}}{\displaystyle \underset{\mu =1}{\overset{M+1}{}}}dx^{(\mu )}\left\{M_{i1}(x^{(1)})p_{t_1}(0,x^{(1)})(1+\chi _1(x^{(1)}))\right\}`$ $`\times {\displaystyle \underset{\mu =1}{\overset{M}{}}}\left\{p_{t_{\mu +1}t_\mu }(x^{(\mu )},x^{(\mu +1)})(1+\chi _{\mu +1}(x^{(\mu +1)}))\right\}p_{Tt_{M+1}}(x^{(M+1)},y_j)],`$ where $`p_{Tt_{M+1}}(x,y)=p_0(x,y)=\delta (xy)`$. For simplicity of expressions, we will assume that the number of particles $`N`$ is even from now on. We use the formula $$_{_<^N}𝑑𝐲\underset{1i,jN}{det}\left[\varphi _i(y_j)\right]=\mathrm{Pf}_{1i,jN}\left[_{}𝑑y_{}𝑑\stackrel{~}{y}\mathrm{sgn}(\stackrel{~}{y}y)\varphi _i(y)\varphi _j(\stackrel{~}{y})\right]$$ for integrable continuous functions $`\varphi _i,1iN`$. Here, for an integer $`n`$ and an antisymmetric $`2n\times 2n`$ matrix $`A=(a_{ij})`$, the pfaffian is defined as $$\mathrm{Pf}(A)=\mathrm{Pf}_{1i<j2n}\left[a_{ij}\right]=\frac{1}{n!}\underset{\sigma }{}\mathrm{sgn}(\sigma )a_{\sigma (1)\sigma (2)}a_{\sigma (3)\sigma (4)}\mathrm{}a_{\sigma (2n1)\sigma (2n)},$$ where the summation is extended over all permutations $`\sigma `$ of $`(1,2,\mathrm{},2n)`$ with restriction $`\sigma (2k1)<\sigma (2k),k=1,2,\mathrm{},n`$. Since $`(\mathrm{Pf}(A))^2=detA`$ for any antisymmetric $`2n\times 2n`$ matrix $`A`$, we have $`\left(Z_{N,T}[\chi ]\right)^2`$ $`=`$ $`\underset{1i,jN}{det}[{\displaystyle }dy{\displaystyle }d\stackrel{~}{y}\mathrm{sgn}(\stackrel{~}{y}y)`$ $`\times {\displaystyle _{^{M+1}}}{\displaystyle \underset{\mu =1}{\overset{M+1}{}}}dx^{(\mu )}\left\{M_{i1}(x^{(1)})p_{t_1}(0,x^{(1)})(1+\chi _1(x^{(1)}))\right\}`$ $`\times {\displaystyle \underset{\mu =1}{\overset{M}{}}}\left\{p_{t_{\mu +1}t_\mu }(x^{(\mu )},x^{(\mu +1)})(1+\chi _{\mu +1}(x^{(\mu +1)}))\right\}p_{Tt_{M+1}}(x^{(M+1)},y)`$ $`\times {\displaystyle _{^{M+1}}}{\displaystyle \underset{\mu =1}{\overset{M+1}{}}}d\stackrel{~}{x}^{(\mu )}\left\{M_{j1}(\stackrel{~}{x}^{(1)})p_{t_1}(0,\stackrel{~}{x}^{(1)})(1+\chi _1(\stackrel{~}{x}^{(1)}))\right\}`$ $`\times {\displaystyle \underset{\mu =1}{\overset{M}{}}}\left\{p_{t_{\mu +1}t_\mu }(\stackrel{~}{x}^{(\mu )},\stackrel{~}{x}^{(\mu +1)})(1+\chi _{\mu +1}(\stackrel{~}{x}^{(\mu +1)}))\right\}p_{Tt_{M+1}}(\stackrel{~}{x}^{(M+1)},\stackrel{~}{y})]`$ $`=`$ $`\underset{1i,jN}{det}\left[(A_0)_{ij}+(A_1)_{ij}+(A_2)_{ij}+(A_3)_{ij}\right]`$ with $`(A_0)_{ij}`$ $`=`$ $`{\displaystyle 𝑑y𝑑\stackrel{~}{y}\mathrm{sgn}(\stackrel{~}{y}y)𝑑xM_{i1}(x)p_{t_1}(0,x)p_{Tt_1}(x,y)}`$ $`\times {\displaystyle }d\stackrel{~}{x}M_{j1}(\stackrel{~}{x})p_{t_1}(0,\stackrel{~}{x})p_{Tt_1}(\stackrel{~}{x},\stackrel{~}{y}),`$ $`(A_1)_{ij}`$ $`=`$ $`{\displaystyle \underset{\mathrm{}=1}{\overset{M+1}{}}}{\displaystyle \underset{1\mu _1<\mu _2<\mathrm{}<\mu _{\mathrm{}}M+1}{}}{\displaystyle 𝑑y𝑑\stackrel{~}{y}\mathrm{sgn}(\stackrel{~}{y}y)}`$ $`\times {\displaystyle _{^{\mathrm{}}}}{\displaystyle \underset{k=1}{\overset{\mathrm{}}{}}}dx^{(\mu _k)}{\displaystyle }dxM_{i1}(x)p_{t_1}(0,x)p_{t_{\mu _1}t_1}(x,x^{(\mu _1)})`$ $`\times {\displaystyle \underset{k=1}{\overset{\mathrm{}1}{}}}\left\{\chi _k(x^{(\mu _k)})p_{t_{\mu _{k+1}}t_{\mu _k}}(x^{(\mu _k)},x^{(\mu _{k+1})})\right\}\chi _{\mathrm{}}(x^{(\mu _{\mathrm{}})})p_{Tt_\mu _{\mathrm{}}}(x^{(\mu _{\mathrm{}})},y)`$ $`\times {\displaystyle }d\stackrel{~}{x}M_{j1}(\stackrel{~}{x})p_{t_1}(0,\stackrel{~}{x})p_{Tt_1}(\stackrel{~}{x},\stackrel{~}{y}),`$ $`(A_2)_{ij}`$ $`=`$ $`{\displaystyle \underset{\mathrm{}=1}{\overset{M+1}{}}}{\displaystyle \underset{1\mu _1<\mu _2<\mathrm{}<\mu _{\mathrm{}}M+1}{}}{\displaystyle 𝑑y𝑑\stackrel{~}{y}\mathrm{sgn}(\stackrel{~}{y}y)}`$ $`\times {\displaystyle }dxM_{i1}(x)p_{t_1}(0,x)p_{Tt_1}(x,y),`$ $`\times {\displaystyle _{^{\mathrm{}}}}{\displaystyle \underset{k=1}{\overset{\mathrm{}}{}}}d\stackrel{~}{x}^{(\mu _k)}{\displaystyle }d\stackrel{~}{x}M_{j1}(\stackrel{~}{x})p_{t_1}(0,\stackrel{~}{x})p_{t_{\mu _1}t_1}(\stackrel{~}{x},\stackrel{~}{x}^{(\mu _1)})`$ $`\times {\displaystyle \underset{k=1}{\overset{\mathrm{}1}{}}}\left\{\chi _k(\stackrel{~}{x}^{(\mu _k)})p_{t_{\mu _{k+1}}t_{\mu _k}}(\stackrel{~}{x}^{(\mu _k)},\stackrel{~}{x}^{(\mu _{k+1})})\right\}\chi _{\mathrm{}}(\stackrel{~}{x}^{(\mu _{\mathrm{}})})p_{Tt_\mu _{\mathrm{}}}(\stackrel{~}{x}^{(\mu _{\mathrm{}})},\stackrel{~}{y}),`$ $`(A_3)_{ij}`$ $`=`$ $`{\displaystyle \underset{\mathrm{}=1}{\overset{M+1}{}}}{\displaystyle \underset{m=1}{\overset{M+1}{}}}{\displaystyle \underset{1\mu _1<\mu _2<\mathrm{}<\mu _{\mathrm{}}M+1}{}}{\displaystyle \underset{1\nu _1<\nu _2<\mathrm{}<\nu _mM+1}{}}{\displaystyle 𝑑y𝑑\stackrel{~}{y}\mathrm{sgn}(\stackrel{~}{y}y)}`$ $`\times {\displaystyle _{^{\mathrm{}}}}{\displaystyle \underset{k=1}{\overset{\mathrm{}}{}}}dx^{(\mu _k)}{\displaystyle }dxM_{i1}(x)p_{t_1}(0,x)p_{t_{\mu _1}t_1}(x,x^{(\mu _1)})`$ $`\times {\displaystyle \underset{k=1}{\overset{\mathrm{}1}{}}}\left\{\chi _k(x^{(\mu _k)})p_{t_{\mu _{k+1}}t_{\mu _k}}(x^{(\mu _k)},x^{(\mu _{k+1})})\right\}\chi _{\mathrm{}}(x^{(\mu _{\mathrm{}})})p_{Tt_\mu _{\mathrm{}}}(x^{(\mu _{\mathrm{}})},y)`$ $`\times {\displaystyle _{^{\mathrm{}}}}{\displaystyle \underset{n=1}{\overset{m}{}}}d\stackrel{~}{x}^{(\nu _m)}{\displaystyle }d\stackrel{~}{x}M_{j1}(\stackrel{~}{x})p_{t_1}(0,\stackrel{~}{x})p_{t_{\nu _1}t_1}(\stackrel{~}{x},\stackrel{~}{x}^{(\nu _1)})`$ $`\times {\displaystyle \underset{n=1}{\overset{m1}{}}}\left\{\chi _n(\stackrel{~}{x}^{(\nu _n)})p_{t_{\nu _{n+1}}t_{\nu _n}}(\stackrel{~}{x}^{(\nu _n)},\stackrel{~}{x}^{(\nu _{n+1})})\right\}\chi _m(\stackrel{~}{x}^{(\nu _m)})p_{Tt_{\nu _m}}(\stackrel{~}{x}^{(\nu _m)},\stackrel{~}{y}),`$ where we have used the Chapman-Kolmogorov equation for the heat-kernel $$𝑑yp_{ts}(x,y)p_{ut}(y,z)=p_{us}(x,z),0<s<t<u,x,y,$$ and the fact that $`p_0(x,y)=lim_{t0}p_t(x,y)=\delta (xy)`$. We consider a vector space $`𝒱`$ with the orthonormal basis $`\left\{|\mu ,x\right\}_{\mu =1,2,\mathrm{},M+1,x}`$ which satisfy $$\mu ,x|\nu ,y=\delta _{\mu \nu }\delta (xy),$$ $`\mu ,\nu =1,2,\mathrm{},M+1,x,y`$. We introduce the operators $`\widehat{J},\widehat{p},\widehat{p}_+,\widehat{p}_{}`$ and $`\widehat{\chi }`$ acting on $`𝒱`$ as follows $`\mu ,x|\widehat{J}|\nu ,y=\mathrm{𝟏}_{(\mu =\nu =M+1)}\mathrm{sgn}(yx),`$ $`\mu ,x|\widehat{p}|\nu ,y=p_{|t_\nu t_\mu |}(x,y),`$ $`\mu ,x|\widehat{p}_+|\nu ,y=p_{t_\nu t_\mu }(x,y)\mathrm{𝟏}_{(\mu <\nu )}=\nu ,y|\widehat{p}_{}|\mu ,x,`$ $`\mu ,x|\widehat{\chi }|\nu ,y=\chi _\mu (x)\delta _{\mu \nu }\delta (xy),`$ where $`\mathrm{𝟏}_{(\omega )}`$ is the indicator function: $`\mathrm{𝟏}_{(\omega )}=1`$ if $`\omega `$ is satisfied and $`\mathrm{𝟏}_{(\omega )}=0`$ otherwise, and we will use the convention $$\mu ,x|\widehat{A}|\nu ,y\nu ,y|\widehat{B}|\rho ,z)=\underset{\nu =1}{\overset{M+1}{}}_{}𝑑yA(\mu ,x;\nu ,y)B(\nu ,y;\rho ,z)=\mu ,x|\widehat{A}\widehat{B}|\rho ,z$$ for operators $`\widehat{A},\widehat{B}`$ with $`\mu ,x|\widehat{A}|\nu ,y=A(\mu ,x;\nu ,y),\mu ,x|\widehat{B}|\nu ,y=B(\mu ,x;\nu ,y)`$. Consider another basis $`\left\{|i;i=1,2,\mathrm{}\right\}`$ in $`𝒱`$ and we assume that transformation matrix between the two bases is given by (2.3) $$i|\mu ,x=\mu ,x|i=𝑑yM_{i1}(y)p_{t_1}(0,y)p_{t_\mu t_1}(y,x),$$ $`i=1,2,\mathrm{},\mu =1,2,\mathrm{},M+1,x`$. Then the quantity $`{\displaystyle 𝑑yM_{i1}(y)p_{t_1}(0,y)p_{t_\mu t_1}(y,x)}`$ $`+{\displaystyle \underset{m=1}{\overset{\mu 1}{}}}{\displaystyle \underset{1\mu _1<\mu _2<\mathrm{}<\mu _m<\mu }{}}{\displaystyle _^m}{\displaystyle \underset{j=1}{\overset{m}{}}}dx^{(\mu _j)}{\displaystyle 𝑑yM_{i1}(y)p_{t_1}(0,y)p_{t_{\mu _1}t_1}(y,x^{(\mu _1)})}`$ $`\times {\displaystyle \underset{k=1}{\overset{m1}{}}}\left\{\chi _{\mu _k}(x^{(\mu _k)})p_{t_{\mu _{k+1}}t_{\mu _k}}(x^{(\mu _k)},x^{(\mu _{k+1})})\right\}\chi _{\mu _m}(x^{(\mu _m)})p_{t_\mu t_{\mu _m}}(x^{(\mu _m)},x),`$ for $`\mu =1,2,\mathrm{},M+1,x,i=1,2,\mathrm{}`$, can be written as $`i|\mu ,x+{\displaystyle \underset{m1}{}}i|\mu _1,x^{(\mu _1)}\mu _1,x^{(\mu _1)}|\widehat{\chi }\widehat{p}_+|\mu _2,x^{(\mu _2)}`$ $`\mathrm{}\mu _{m1},x^{(\mu _{m1})}|\widehat{\chi }\widehat{p}_+|\mu _m,x^{(\mu _m)}\mu _m,x^{(\mu _m)}|\widehat{\chi }\widehat{p}_+|\mu ,x`$ $`=`$ $`i|\mu ,x+{\displaystyle \underset{m1}{}}i|(\widehat{\chi }\widehat{p}_+)^m|\mu ,x`$ $`=`$ $`i|{\displaystyle \frac{1}{1\widehat{\chi }\widehat{p}_+}}|\mu ,x.`$ It is also expressed as $`\mu ,x|{\displaystyle \frac{1}{1\widehat{p}_{}\widehat{\chi }}}|i.`$ It should be noted that the basis $`\left\{|i;i=1,2,\mathrm{}\right\}`$ is in general not orthonormal. Here we introduce an operator $`\widehat{\delta }`$ such that (2.4) $$i|\widehat{\delta }|j=j|\widehat{\delta }|i=\delta _{ij},i,j=1,2,\mathrm{}.$$ We will use the following convention $$i|\widehat{A}|jj|\widehat{B}|\mu ,x=\underset{j=1}{\overset{\mathrm{}}{}}A_{ij}B_j^{(\mu )}(x)$$ for $`A_{ij}=i|\widehat{A}|ji|\mu ,x\mu ,x|\widehat{A}|\nu ,y\nu ,y|j,B_j^{(\mu )}(x)=j|\widehat{B}|\mu ,xj|\nu ,y\nu ,y|\widehat{B}|\mu ,x,`$ but we will not write it as $`i|\widehat{A}\widehat{B}|\mu ,x`$, since $`\left\{|i;i=1,2,\mathrm{}\right\}`$ is in general not a complete basis. By this basis, any operator $`\widehat{A}`$ on $`𝒱`$ may have a semi-infinite matrix representation $`A=\left(i|\widehat{A}|j\right)_{i,j=1,2,\mathrm{}}`$. If the matrix $`A`$ representing an operator $`\widehat{A}`$ is invertible, we define the operator $`\widehat{A}^{}`$ such that its matrix representation is the inverse of $`A`$; (2.5) $$\left(i|\widehat{A}^{}|j\right)_{i,j=1,2,\mathrm{}}=A^1.$$ In other words, $$i|\widehat{A}|jj|\widehat{A}^{}|k=i|\widehat{A}^{}|jj|\widehat{A}|k=i|\widehat{\delta }|k,i,k=1,2,\mathrm{}.$$ For any given operator $`\widehat{A}`$, the equality $$i|\widehat{A}|jj|\widehat{A}^{}|kk|\widehat{B}|\mathrm{}=i|\widehat{B}|\mathrm{}.$$ holds for arbitrary $`i,\mathrm{}=1,2,\mathrm{}`$ and $`\widehat{B}`$. Then the equality (2.6) $$\widehat{A}|jj|\widehat{A}^{}|kk|=1$$ should be established for each $`\widehat{A}`$. We will use this equality later. Let $`𝒫_N`$ be a projection operator from the space spanned by $`\left\{|i;i=1,2,\mathrm{}\right\}`$ to its $`N`$-dimensional subspace spanned by $`\left\{|i;i=1,2,\mathrm{},N\right\}`$, and thus $$i|𝒫_N|\mu ,x=\mu ,x|𝒫_N|i=\{\begin{array}{cc}i|\mu ,x,\hfill & \text{if}1iN,\hfill \\ & \\ 0,\hfill & \text{otherwise}.\hfill \end{array}$$ Then we have the following expressions for $`(A_\alpha )_{ij},\alpha =0,1,2,3,i,j=1,2,\mathrm{},N`$, $`(A_0)_{ij}`$ $`=`$ $`i|𝒫_N|\mu ,x\mu ,x|\widehat{J}|\nu ,y\nu ,y|𝒫_N|j`$ $`=`$ $`i|𝒫_N\widehat{J}𝒫_N|j,`$ $`(A_1)_{ij}`$ $`=`$ $`i|𝒫_N{\displaystyle \frac{1}{1\widehat{\chi }\widehat{p}_+}}\widehat{\chi }|\mu ,x\mu ,x|\widehat{p}\widehat{J}𝒫_N|j`$ $`=`$ $`i|𝒫_N\widehat{\chi }{\displaystyle \frac{1}{1\widehat{p}_+\widehat{\chi }}}|\mu ,x\mu ,x|\widehat{p}\widehat{J}𝒫_N|j`$ $`=`$ $`i|𝒫_N\widehat{\chi }{\displaystyle \frac{1}{1\widehat{p}_+\widehat{\chi }}}\widehat{p}\widehat{J}𝒫_N|j,`$ $`(A_2)_{ij}`$ $`=`$ $`i|𝒫_N\widehat{J}\widehat{p}|\mu ,x\mu ,x|\widehat{\chi }{\displaystyle \frac{1}{1\widehat{p}_{}\widehat{\chi }}}𝒫_N|j,`$ $`=`$ $`i|𝒫_N\widehat{J}\widehat{p}\widehat{\chi }{\displaystyle \frac{1}{1\widehat{p}_{}\widehat{\chi }}}𝒫_N|j,`$ $`(A_3)_{ij}`$ $`=`$ $`i|𝒫_N{\displaystyle \frac{1}{1\widehat{\chi }\widehat{p}_+}}\widehat{\chi }|\mu ,x\mu ,x|\widehat{p}\widehat{J}\widehat{p}|\nu ,y\nu ,y|\widehat{\chi }{\displaystyle \frac{1}{1\widehat{p}_{}\widehat{\chi }}}𝒫_N|j`$ $`=`$ $`i|𝒫_N\widehat{\chi }{\displaystyle \frac{1}{1\widehat{p}_+\widehat{\chi }}}|\mu ,x\mu ,x|\widehat{p}\widehat{J}\widehat{p}|\nu ,y\nu ,y|\widehat{\chi }{\displaystyle \frac{1}{1\widehat{p}_{}\widehat{\chi }}}𝒫_N|j`$ $`=`$ $`i|𝒫_N\widehat{\chi }{\displaystyle \frac{1}{1\widehat{p}_+\widehat{\chi }}}\widehat{p}\widehat{J}\widehat{p}\widehat{\chi }{\displaystyle \frac{1}{1\widehat{p}_{}\widehat{\chi }}}𝒫_N|j.`$ That is, the matrices $`A_\alpha =\left((A_\alpha )_{ij}\right)_{i,j=1,2,\mathrm{},N},\alpha =0,1,2,3,`$ can be regarded as the matrix representations in the basis $`\left\{|i;i=1,2,\mathrm{}\right\}`$ of the operators. Since $`\left(Z_{N,T}[0]\right)^2=det_{1i,jN}\left[(A_0)_{ij}\right]`$, (2.1) gives (2.7) $$\left\{\mathrm{\Psi }_{N,T}(𝐟;𝜽)\right\}^2=\underset{1i,jN}{det}\left[\delta _{ij}+(A_0^1A_1)_{ij}+(A_0^1A_2)_{ij}+(A_0^1A_3)_{ij}\right].$$ By our notation (2.5), $$(A_0^1)_{ij}=i|𝒫_N(𝒫_N\widehat{J}𝒫_N)^{}𝒫_N|j,$$ since it satisfies the relation $`i|𝒫_N(𝒫_N\widehat{J}𝒫_N)^{}𝒫_N|jj|𝒫_N\widehat{J}𝒫_N|k`$ $`=`$ $`\mathrm{𝟏}_{(1i,kN)}{\displaystyle \underset{j=1}{\overset{N}{}}}(A_0^1)_{ij}(A_0)_{jk}=\mathrm{𝟏}_{(1i,kN)}\delta _{ik}`$ $`=`$ $`i|𝒫_N\widehat{\delta }𝒫_N|j.`$ We will use the abbreviation $$\widehat{A}_N=𝒫_N\widehat{A}𝒫_N$$ for an operator $`\widehat{A}`$. Then it is easy to see that (2.7) is written in the form $`\left\{\mathrm{\Psi }_{N,T}(𝐟;𝜽)\right\}^2`$ $`=`$ $`\underset{1i,jN}{det}\left[\delta _{ij}+i|𝐁|\mu ,x\mu ,x|𝐂|j\right]`$ $`=`$ $`\underset{1i,jN}{det}i|\left[\widehat{\delta }_N+\mathrm{𝐁𝐂}\right]|j,`$ where $`i|𝐁|\mu ,x`$ and $`\mu ,x|𝐂|j`$ are the two-dimensional row and column vectors, respectively, given by (2.12) $`i\left|𝐁\right|\mu ,x`$ $`=`$ $`\left(\begin{array}{ccc}i\left|𝒫_N\widehat{J}_N^{}𝒫_N\right|kk\left|\widehat{\chi }\frac{1}{1\widehat{p}_+\widehat{\chi }}\right|\mu ,x\hfill & & i\left|𝒫_N\widehat{J}_N^{}𝒫_N\right|kk\left|\widehat{J}\widehat{p}\widehat{\chi }\right|\mu ,x\hfill \\ & & i\left|𝒫_N\widehat{J}_N^{}𝒫_N\right|k\hfill \\ & & \times k\left|\widehat{\chi }\frac{1}{1\widehat{p}_+\widehat{\chi }}\right|\nu ,y\nu ,y\left|\widehat{p}\widehat{J}\widehat{p}\widehat{\chi }\right|\mu ,x\hfill \end{array}\right),`$ $`\mu ,x\left|𝐂\right|j`$ $`=`$ $`\left(\begin{array}{c}\mu ,x\left|\widehat{p}\widehat{J}𝒫_N\right|j\\ \\ \mu ,x\left|\frac{1}{1\widehat{p}_{}\widehat{\chi }}𝒫_N\right|j\end{array}\right).`$ The determinant (2) is written using the Fredholm determinant, (2.21) $`\mathrm{Det}\mu ,x\left|\left[I_2+\mathrm{𝐂𝐁}\right]\right|\nu ,y`$ $`=`$ $`\mathrm{Det}\left[\begin{array}{cc}\mu ,x|\nu ,y\hfill & \mu ,x\left|\widehat{p}\widehat{J}\right|ii\left|𝒫_N\widehat{J}_N^{}𝒫_N\right|jj\left|\widehat{J}\widehat{p}\widehat{\chi }\right|\nu ,y\hfill \\ +\mu ,x\left|\widehat{p}\widehat{J}\right|i\hfill & \mu ,x\left|\widehat{p}\widehat{J}\right|ii\left|𝒫_N\widehat{J}_N^{}𝒫_N\right|jj\left|\widehat{\chi }\frac{1}{1\widehat{p}_+\widehat{\chi }}\right|\rho ,z\hfill \\ \times i\left|𝒫_N\widehat{J}_N^{}𝒫_N\right|jj\left|\widehat{\chi }\frac{1}{1\widehat{p}_+\widehat{\chi }}\right|\nu ,y\hfill & \times \rho ,z\left|\widehat{p}\widehat{J}\widehat{p}\widehat{\chi }\right|\nu ,y\hfill \\ & \\ \mu ,x\left|\frac{1}{1\widehat{p}_{}\widehat{\chi }}\right|i\hfill & \mu ,x|\nu ,y\hfill \\ \times i\left|𝒫_N\widehat{J}_N^{}𝒫_N\right|jj\left|\widehat{\chi }\frac{1}{1\widehat{p}_+\widehat{\chi }}\right|\nu ,y\hfill & +\mu ,x\left|\frac{1}{1\widehat{p}_{}\widehat{\chi }}\right|ii\left|𝒫_N\widehat{J}_N^{}𝒫_N\right|jj\left|\widehat{J}\widehat{p}\widehat{\chi }\right|\nu ,y\hfill \\ & +\mu ,x\left|\frac{1}{1\widehat{p}_{}\widehat{\chi }}\right|ii\left|𝒫_N\widehat{J}_N^{}𝒫_N\right|jj\left|\widehat{\chi }\frac{1}{1\widehat{p}_+\widehat{\chi }}\right|\rho ,z\hfill \\ & \times \rho ,z\left|\widehat{p}\widehat{J}\widehat{p}\widehat{\chi }\right|\nu ,y\hfill \end{array}\right]`$ $`=`$ $`\mathrm{Det}(\left[\begin{array}{cc}\mu ,x|\rho ,z\hfill & \mu ,x\left|\widehat{p}\widehat{J}\right|i\hfill \\ +\mu ,x\left|\widehat{p}\widehat{J}\right|i\hfill & \times i\left|𝒫_N\widehat{J}_N^{}𝒫_N\right|jj\left|\widehat{J}\widehat{p}\widehat{\chi }\right|\rho ,z\hfill \\ \times i\left|𝒫_N\widehat{J}_N^{}𝒫_N\right|jj\left|\widehat{\chi }\frac{1}{1\widehat{p}_+\widehat{\chi }}\right|\rho ,z\hfill & +\mu ,x\left|\widehat{p}\widehat{J}\widehat{p}\widehat{\chi }\right|\rho ,z\hfill \\ & \\ \mu ,x\left|\frac{1}{1\widehat{p}_{}\widehat{\chi }}\right|i\hfill & \mu ,x|\rho ,z\hfill \\ \times i\left|𝒫_N\widehat{J}_N^{}𝒫_N\right|jj\left|\widehat{\chi }\frac{1}{1\widehat{p}_+\widehat{\chi }}\right|\rho ,z\hfill & +\mu ,x\left|\frac{1}{1\widehat{p}_{}\widehat{\chi }}\right|i\hfill \\ & \times i\left|𝒫_N\widehat{J}_N^{}𝒫_N\right|jj\left|\widehat{J}\widehat{p}\widehat{\chi }\right|\rho ,z\hfill \end{array}\right]`$ $`\times \left[\begin{array}{cc}& \\ \rho ,z|\nu ,y& \rho ,z\left|\widehat{p}\widehat{J}\widehat{p}\widehat{\chi }\right|\nu ,y\\ & \\ & \\ 0& \rho ,z|\nu ,y\end{array}\right])`$ (2.37) $`=`$ $`\mathrm{Det}\left[\begin{array}{cc}\mu ,x|\nu ,y\hfill & \mu ,x\left|\widehat{p}\widehat{J}\right|i\hfill \\ +\mu ,x\left|\widehat{p}\widehat{J}\right|i\hfill & \times i\left|𝒫_N\widehat{J}_N^{}𝒫_N\right|jj\left|\widehat{J}\widehat{p}\widehat{\chi }\right|\nu ,y\hfill \\ \times i\left|𝒫_N\widehat{J}_N^{}𝒫_N\right|jj\left|\widehat{\chi }\frac{1}{1\widehat{p}_+\widehat{\chi }}\right|\nu ,y\hfill & +\mu ,x\left|\widehat{p}\widehat{J}\widehat{p}\widehat{\chi }\right|\nu ,y\hfill \\ & \\ \mu ,x\left|\frac{1}{1\widehat{p}_{}\widehat{\chi }}\right|i\hfill & \mu ,x|\nu ,y\hfill \\ \times i\left|𝒫_N\widehat{J}_N^{}𝒫_N\right|jj\left|\widehat{\chi }\frac{1}{1\widehat{p}_+\widehat{\chi }}\right|\nu ,y\hfill & +\mu ,x\left|\frac{1}{1\widehat{p}_{}\widehat{\chi }}\right|i\hfill \\ & \times i\left|𝒫_N\widehat{J}_N^{}𝒫_N\right|jj\left|\widehat{J}\widehat{p}\widehat{\chi }\right|\nu ,y\hfill \end{array}\right],`$ where $`I_2`$ denotes the unit matrix with size 2. It is further rewritten as $`\mathrm{Det}\mu ,x|(I_2+\left[\begin{array}{cc}1& 0\\ & \\ 0& \frac{1}{1\widehat{p}_{}\widehat{\chi }}\end{array}\right]`$ $`\times \left[\begin{array}{cc}\widehat{p}\widehat{J}|ii\left|𝒫_N\widehat{J}_N^{}𝒫_N\right|jj|\widehat{\chi }& \left(\widehat{p}\widehat{J}|ii\left|𝒫_N\widehat{J}_N^{}𝒫_N\right|jj|\widehat{J}\widehat{p}+\widehat{p}\widehat{J}\widehat{p}\right)\widehat{\chi }\\ & \\ |ii\left|𝒫_N\widehat{J}_N^{}𝒫_N\right|jj|\widehat{\chi }& |ii\left|𝒫_N\widehat{J}_N^{}𝒫_N\right|jj|\widehat{J}\widehat{p}\widehat{\chi }\end{array}\right]`$ $`\times \left[\begin{array}{cc}\frac{1}{1\widehat{p}_+\widehat{\chi }}& 0\\ & \\ 0& 1\end{array}\right])|\nu ,y`$ $`=`$ $`\mathrm{Det}\mu ,x|(\left[\begin{array}{cc}1& 0\\ & \\ 0& 1\widehat{p}_{}\widehat{\chi }\end{array}\right]\left[\begin{array}{cc}1\widehat{p}_+\widehat{\chi }& 0\\ & \\ 0& 1\end{array}\right]`$ $`+\left[\begin{array}{cc}\widehat{p}\widehat{J}|ii\left|𝒫_N\widehat{J}_N^{}𝒫_N\right|jj|\widehat{\chi }& \left(\widehat{p}\widehat{J}|ii\left|𝒫_N\widehat{J}_N^{}𝒫_N\right|jj|\widehat{J}\widehat{p}+\widehat{p}\widehat{J}\widehat{p}\right)\widehat{\chi }\\ & \\ |ii\left|𝒫_N\widehat{J}_N^{}𝒫_N\right|jj|\widehat{\chi }& |ii\left|𝒫_N\widehat{J}_N^{}𝒫_N\right|jj|\widehat{J}\widehat{p}\widehat{\chi }\end{array}\right])|\nu ,y`$ $`=`$ $`\mathrm{Det}\mu ,x\left|\left(I_2+\left[\begin{array}{cc}\widehat{p}\widehat{J}|ii\left|𝒫_N\widehat{J}_N^{}𝒫_N\right|jj|\widehat{p}_+& \widehat{p}\widehat{J}|ii\left|𝒫_N\widehat{J}_N^{}𝒫_N\right|jj|\widehat{J}\widehat{p}+\widehat{p}\widehat{J}\widehat{p}\\ & \\ |ii\left|𝒫_N\widehat{J}_N^{}𝒫_N\right|jj|& |ii\left|𝒫_N\widehat{J}_N^{}𝒫_N\right|jj|\widehat{J}\widehat{p}\widehat{p}_{}\end{array}\right]\widehat{\chi }\right)\right|\nu ,y.`$ Here we have used the facts that $$\mathrm{Det}\mu ,x|\left[\begin{array}{cc}1& 0\\ & \\ 0& 1\widehat{p}_{}\widehat{\chi }\end{array}\right]|\nu ,y=1,$$ and $$\mathrm{Det}\mu ,x|\left[\begin{array}{cc}1\widehat{p}_+\widehat{\chi }& 0\\ & \\ 0& 1\end{array}\right]|\nu ,y=1,$$ which are consequences of definitions (2) of the operators $`\widehat{p}_+`$ and $`\widehat{p}_{}`$. Then we arrive at (2.38) $$\left\{\mathrm{\Psi }_{N,T}(𝐟;𝜽)\right\}^2=\mathrm{Det}\left(I_2\delta _{\mu \nu }\delta (xy)+\left[\begin{array}{cc}\stackrel{~}{S}^{\mu ,\nu }(x,y)& \stackrel{~}{I}^{\mu ,\nu }(x,y)\\ D^{\mu ,\nu }(x,y)& \stackrel{~}{S}^{\nu ,\mu }(y,x)\end{array}\right]\chi _\nu (y)\right),$$ where (2.39) $`D^{\mu ,\nu }(x,y)`$ $`=`$ $`\mu ,x|ii|𝒫_N(𝒫_N\widehat{J}𝒫_N)^{}𝒫_N|jj|\nu ,y,`$ $`S^{\mu ,\nu }(x,y)`$ $`=`$ $`\mu ,x|\widehat{p}\widehat{J}|ii|𝒫_N(𝒫_N\widehat{J}𝒫_N)^{}𝒫_N|jj|\nu ,y,`$ $`I^{\mu ,\nu }(x,y)`$ $`=`$ $`\mu ,x|\widehat{p}\widehat{J}|ii|𝒫_N(𝒫_N\widehat{J}𝒫_N)^{}𝒫_N|jj|\widehat{J}\widehat{p}|\nu ,y,`$ and (2.44) $`\stackrel{~}{S}^{\mu ,\nu }(x,y)`$ $`=`$ $`S^{\mu ,\nu }(x,y)\mu ,x|\widehat{p}_+|\nu ,y`$ $`=`$ $`\{\begin{array}{cc}\mu ,x|\widehat{p}\widehat{J}|ii|𝒫_N(𝒫_N\widehat{J}𝒫_N)^{}𝒫_N|jj|\nu ,y,\hfill & \text{if }\mu \nu \text{,}\hfill \\ & \\ \mu ,x|\widehat{p}\widehat{J}|ii|\left(\widehat{J}^{}𝒫_N(𝒫_N\widehat{J}𝒫_N)^{}𝒫_N\right)|jj|\nu ,y,\hfill & \text{if }\mu <\nu \text{,}\hfill \end{array}`$ $`\stackrel{~}{I}^{\mu ,\nu }(x,y)`$ $`=`$ $`I^{\mu ,\nu }(x,y)+\mu ,x|\widehat{p}\widehat{J}\widehat{p}|\nu ,y`$ $`=`$ $`\mu ,x|\widehat{p}\widehat{J}|ii|\left(\widehat{J}^{}𝒫_N(𝒫_N\widehat{J}𝒫_N)^{}𝒫_N\right)|jj|\widehat{J}\widehat{p}|\nu ,y,`$ where we have used the equality (2.6) for $`\widehat{A}=\widehat{J}`$. In Rains introduced Fredholm pfaffian, denoted here by $`\mathrm{PF}`$, and proved a useful equality (2.45) $$\left\{\mathrm{PF}(J_2+K)\right\}^2=\mathrm{Det}(I_2+J_2^1K),$$ for any antisymmetric $`2\times 2`$ matrix kernel $`K`$, where (2.46) $$J_2=\left(\begin{array}{cc}0& 1\\ 1& 0\end{array}\right).$$ Then (2.38) implies the Fredholm pfaffian representation of the multi-time characteristic function (2.47) $$\mathrm{\Psi }_{N,T}(𝐟;𝜽)=\mathrm{PF}\left(J_2\delta _{\mu \nu }\delta (xy)+\left[\begin{array}{cc}D^{\mu ,\nu }(x,y)& \stackrel{~}{S}^{\nu ,\mu }(y,x)\\ \stackrel{~}{S}^{\mu ,\nu }(x,y)& \stackrel{~}{I}^{\mu ,\nu }(x,y)\end{array}\right]\chi _\nu (y)\right).$$ ## 3 Pfaffian Process Let (3.1) $$A^{\mu ,\nu }(x,y)=\left[\begin{array}{cc}D^{\mu ,\nu }(x,y)& \stackrel{~}{S}^{\nu ,\mu }(y,x)\\ \stackrel{~}{S}^{\mu ,\nu }(x,y)& \stackrel{~}{I}^{\mu ,\nu }(x,y)\end{array}\right],$$ and construct $`2_{\mu =1}^{M+1}N_\mu \times 2_{\mu =1}^{M+1}N_\mu `$ antisymmetric matrices $$A(𝐱_{N_1}^{(1)},𝐱_{N_2}^{(2)},\mathrm{},𝐱_{N_{M+1}}^{(M+1)})=\left(A^{\mu ,\nu }(x_i^{(\mu )},x_j^{(\nu )})\right)_{1iN_\mu ,1jN_\nu ,1\mu ,\nu M+1},$$ for $`N_m=1,2,\mathrm{},N,1mM+1`$. By the definition of Fredholm pfaffian and the equality (2.47), we can establish the following statement. ###### Theorem 1 The non-colliding system of Brownian motions $`𝐗(t)`$ is a pfaffian process in the sense that any multi-time correlation function is given by a pfaffian (3.2) $`\rho _{N,T}(t_1,\{𝐱_{N_1}^{(1)}\};t_2,\{𝐱_{N_2}^{(2)}\};\mathrm{};t_{M+1},\{𝐱_{N_{M+1}}^{(M+1)}\})`$ (3.3) $`=\mathrm{Pf}\left[A(𝐱_{N_1}^{(1)},𝐱_{N_2}^{(2)},\mathrm{},𝐱_{N_{M+1}}^{(M+1)})\right].`$ Remark. The pfaffian processes considered here may be regarded as the continuous space-time version of the pfaffian point processes and pfaffian Schur processes studied by Sasamoto and Imamura and by Borodin and Rains . The processes studied in are also pfaffian processes, since the ‘quaternion determinantal expressions’, in the sense of Dyson and Mehta , of correlation functions are readily transformed to pfaffian expressions. Let $`H_i(x)`$ be the $`i`$-th Hermite polynomial $`H_i(x)`$ $`=`$ $`e^{x^2}\left({\displaystyle \frac{d}{dx}}\right)^ie^{x^2}`$ $`=`$ $`i!{\displaystyle \underset{j=0}{\overset{[i/2]}{}}}(1)^j{\displaystyle \frac{(2x)^{i2j}}{j!(i2j)!}},`$ where $`[a]`$ denotes the greatest integer not greater than $`a`$. The Hermite polynomials satisfy the orthogonal relations (3.4) $$_{}𝑑xe^{x^2}H_i(x)H_j(x)=2^ii!\sqrt{\pi }\delta _{ij},i,j=0,1,2,\mathrm{}.$$ Set $$c_1=\sqrt{\frac{t_1(2Tt_1)}{T}},z_1=\sqrt{\frac{2Tt_1}{t_1}},$$ and (3.5) $$\alpha _{ij}=\{\begin{array}{cc}2^ic_1^i\delta _{ij},\hfill & \text{if }i\text{ is even,}\hfill \\ 2^ic_1^i\left\{\delta _{ij}2(i1)\delta _{i2j}\right\},\hfill & \text{if }i\text{ is odd.}\hfill \end{array}$$ Now we specify polynomials $`\{M_i(x)\}`$ as (3.6) $$M_i(x)=b_iz_1^i\underset{j=0}{\overset{i}{}}\alpha _{ij}H_j\left(\frac{x}{c_1}\right)z_1^j,i=0,1,2,\mathrm{}$$ with $`b_i=\left\{r_{[i/2]}\right\}^{1/2}`$, where $`r_i={\displaystyle \frac{1}{\pi }}\mathrm{\Gamma }(i+1/2)\mathrm{\Gamma }(i+1)\left({\displaystyle \frac{t_1^2}{T}}\right)^{2i+1/2}.`$ Set $$J_N=I_{N/2}J_2,$$ where $`I_{N/2}`$ denotes the unit matrix with size $`N/2`$ and $`J_2`$ is given by (2.46), and let $`J`$ be the semi-infinite matrix obtained as the $`N\mathrm{}`$ limit of $`J_N`$. By the orthogonality of Hermite polynomials (3.4), we can show through (2.3) with the choice (3.6) that $$i|\widehat{J}|j=i|\mu ,x\mu ,x|\widehat{J}|\nu ,y\nu ,y|j=J_{ij},i,j=1,2,\mathrm{}.$$ Since $`J_N^2=I_N`$ for any even $`N2`$, this implies that the matrix $`\left(i|\widehat{J}|j\right)_{i,j=1,2,\mathrm{}}`$ is invertible and (3.7) $`i|\widehat{J}^{}|j`$ $`=`$ $`=J_{ij},`$ (3.8) $`i|𝒫_N(𝒫_N\widehat{J}𝒫_N)^{}𝒫_N|j`$ $`=`$ $`i|𝒫_N\widehat{J}^{}𝒫_N|j`$ (3.11) $`=`$ $`\{\begin{array}{cc}J_{ij}=(J_N)_{ij},\hfill & \text{if }1i,jN\text{,}\hfill \\ 0,\hfill & \text{otherwise}\hfill \end{array}`$ for $`i,j=1,2,\mathrm{}.`$ If we write $`\mu ,x|i=b_iR_{i1}^{(\mu )}(x),`$ $`i|\widehat{J}\widehat{p}|\mu ,x=\mu ,x|\widehat{p}\widehat{J}|i=b_i\mathrm{\Phi }_{i1}^{(\mu )}(x),`$ $`i=1,2,\mathrm{},\mu =1,2,\mathrm{},M+1,x`$, then the functions (2.39) are written as (3.12) $`D^{\mu ,\nu }(x,y)`$ $`=`$ $`{\displaystyle \underset{i=0}{\overset{N/21}{}}}{\displaystyle \frac{1}{r_i}}\left[R_{2i}^{(\mu )}(x)R_{2i+1}^{(\nu )}(y)R_{2i+1}^{(\mu )}(x)R_{2i}^{(\nu )}(y)\right],`$ $`S^{\mu ,\nu }(x,y)`$ $`=`$ $`{\displaystyle \underset{i=0}{\overset{N/21}{}}}{\displaystyle \frac{1}{r_i}}\left[\mathrm{\Phi }_{2i}^{(\mu )}(x)R_{2i+1}^{(\nu )}(y)\mathrm{\Phi }_{2i+1}^{(\mu )}(x)R_{2i}^{(\nu )}(y)\right],`$ $`I^{\mu ,\nu }(x,y)`$ $`=`$ $`{\displaystyle \underset{i=0}{\overset{N/21}{}}}{\displaystyle \frac{1}{r_i}}\left[\mathrm{\Phi }_{2i}^{(\mu )}(x)\mathrm{\Phi }_{2i+1}^{(\nu )}(y)\mathrm{\Phi }_{2i+1}^{(\mu )}(x)\mathrm{\Phi }_{2i}^{(\nu )}(y)\right],`$ and Equations (LABEL:eqn:DSI2) become (3.16) $`\stackrel{~}{S}^{\mu ,\nu }(x,y)`$ $`=`$ $`\{\begin{array}{cc}{\displaystyle \underset{i=0}{\overset{N/21}{}}}{\displaystyle \frac{1}{r_i}}\left[\mathrm{\Phi }_{2i}^{(\mu )}(x)R_{2i+1}^{(\nu )}(y)\mathrm{\Phi }_{2i+1}^{(\mu )}(x)R_{2i}^{(\nu )}(y)\right],\hfill & \text{if }\mu \nu \text{,}\hfill \\ & \\ {\displaystyle \underset{i=N/2}{\overset{\mathrm{}}{}}}{\displaystyle \frac{1}{r_i}}\left[\mathrm{\Phi }_{2i}^{(\mu )}(x)R_{2i+1}^{(\nu )}(y)\mathrm{\Phi }_{2i+1}^{(\mu )}(x)R_{2i}^{(\nu )}(y)\right],\hfill & \text{if }\mu <\nu \text{,}\hfill \end{array}`$ $`\stackrel{~}{I}^{\mu ,\nu }(x,y)`$ $`=`$ $`{\displaystyle \underset{i=N/2}{\overset{\mathrm{}}{}}}{\displaystyle \frac{1}{r_i}}\left[\mathrm{\Phi }_{2i}^{(\mu )}(x)\mathrm{\Phi }_{2i+1}^{(\nu )}(y)\mathrm{\Phi }_{2i+1}^{(\mu )}(x)\mathrm{\Phi }_{2i}^{(\nu )}(y)\right].`$ Theorem 1 with the expressions (3.12), (3.16) of the elements of matrix (3.1) is equivalent with Theorem 3 reported in , although the latter was given in the form of quaternion determinant. The present argument will be generalized to discuss other non-colliding systems of diffusion particles reported in . ## Acknowledgment This manuscript is based on the joint work with H. Tanemura. The author would like to thank T. Sasamoto for useful discussions on the papers and .
warning/0506/cond-mat0506725.html
ar5iv
text
# Entropy Optimization of Scale-Free Networks Robustness to Random Failures ## 1 Introduction Many complex systems are characterized by the network of interactions among its components. It has been shown that most of the networks are highly heterogeneous in their connectivity patterns. Heterogeneity can be easily identified by looking at the degree distribution $`p(k)`$, which gives the probability of having a node with $`k`$ links. Most complex networks can be described by degree distributions $`p(k)ck^\alpha `$, with $`\alpha (2,3)`$. These networks include social networks(such as movie-actor networks, science citations and cooperation networks), the Internet and the World Wide Web, metabolic networks and protein interaction networks, etc . Since Albert *et al.* raised the question of random failures and intentional attack on networks , enormous interest has been devoted to the study of the resilience of networks to failures of nodes or to intentional attacks . Magoni studied a general strategy of attacks on the Internet . It is important to understand how to design networks which are optimally robust against both failures and attacks. Most of researchers use percolation theory to study this problem . A fraction $`p`$ of the nodes and their connections are removed randomly$``$the integrity might be compromised: for $`\alpha >3`$, there exists a certain threshold value $`p_c`$. When $`p>p_c`$, the network disintegrates into smaller, disconnected parts. Below that critical threshold value, there still exists a connected cluster. For $`2<\alpha <3`$ , the networks are more resilient and the threshold value $`p_c`$ approaches 1 . Also, some researchers work on optimizing network robustness to both random failures and attacks with percolation theory . A simple but essential character of scale-free network is its heterogeneous link distribution. Moreover, heterogeneity is in direct relationship with the network’s resilience to attacks. Many real-world networks are scale-free and they are robust to random failures but vulnerable to targeted attacks. Heterogeneity can be measured by entropy . Solé *et al.* used entropy of the remaining degree and mutual information to study some model networks of different heterogeneity and randomness . In this paper, different from percolation theory, we try an alternative point of view$``$the entropy of the degree distribution to describe scale-free network’s heterogeneity. To determine the optimal design of scale-free network to random failures, we maximize the robustness of scale-free network to random failures while keeping the cost constant (that is, the average number of links per node remains constant). We find that optimization of scale-free network’s robustness to random failures is equivalent to maximize the entropy of the degree distribution. By optimizing the entropy of the degree distribution, we get the scenario of optimal design of scale-free network to random failures. ## 2 Entropy of the degree distribution In general, scale-free networks have the degree distributions $`p(k)`$, $`p(k)ck^\alpha `$, $`k=m,m+1,\mathrm{},K`$, where $`m`$ is the minimal connectivity and $`K`$ is an effective connectivity cutoff present in finite networks. Then the entropy of the degree distribution can be defined as follows: $$H=\underset{k=1}{\overset{N1}{}}p(k)\mathrm{log}p(k)$$ (1) where $`N`$ is the total number of nodes in the network and the restriction to the degree distribution is $`p(k)ck^\alpha `$. Within the context of complex networks, the entropy of the degree distribution provides an average measure of network’s heterogeneity, since it measures the diversity of the link distribution . Two extreme cases are the maximal value and the minimal one. The maximum value is $`H_{max}=\mathrm{log}(N1)`$ obtained for $`p(k)=\frac{1}{N1}(k=1,2,\mathrm{},N1)`$. $`H_{min}=0`$ occurs when $`p(k)=\{0,\mathrm{},1,\mathrm{},0\}`$. For the power-law degree distribution of scale-free network, the impact of diversity (long tails) is obvious, increasing the uncertainty. As the scaling exponent increases or the cut-off decreases, the network becomes less heterogeneous and as a result a lower entropy is observed . In this paper, we study the entropy of the degree distribution for scale-free networks. With continuous approximation, for the power law degree distribution of scale-free network, entropy $`H`$ can be expressed as follows: $`H`$ $`=`$ $`{\displaystyle _m^K}p(k)\mathrm{log}p(k)𝑑k`$ (2) $`=`$ $`{\displaystyle _m^K}ck^\alpha \mathrm{log}(ck^\alpha )𝑑k`$ $`=`$ $`{\displaystyle \frac{c\mathrm{log}c}{1\alpha }}(K^{1\alpha }m^{1\alpha })+{\displaystyle \frac{c\alpha }{1\alpha }}[\mathrm{log}KK^{1\alpha }\mathrm{log}mm^{(1\alpha )}`$ $`{\displaystyle \frac{1}{1\alpha }}(K^{1\alpha }m^{1\alpha })]`$ The largest connectivity $`KmN^{\frac{1}{\alpha 1}}`$ , The coefficient $`c`$ is obtained from $`1`$ $`=`$ $`{\displaystyle _m^{\mathrm{}}}p(k)𝑑k={\displaystyle _m^{\mathrm{}}}ck^\alpha 𝑑k={\displaystyle \frac{c}{1\alpha }}(m^{1\alpha })`$ and thus $`c=(\alpha 1)m^{\alpha 1}`$ (4) Then entropy $`H`$ can be expressed as a function of scaling exponent $`\alpha `$, the minimal connectivity $`m`$ and network scale $`N`$. $$H(\alpha ,m,N)=(\mathrm{log}(\frac{\alpha 1}{m})+\frac{\alpha }{1\alpha })(\frac{1}{N}1)\frac{\alpha }{\alpha 1}\frac{\mathrm{log}N}{N}$$ (5) Fig. 1 displays the relationship between entropy $`H`$ and $`\alpha `$ for different $`N`$ when the minimal connectivity $`m`$ is defined. We know that with the increase of scaling exponent $`\alpha `$, the cut-off connectivity $`K`$ decreases, which makes the network less heterogeneous. As a result, a lower entropy is observed. So different scaling exponent $`\alpha `$ brings the network different degree of heterogeneity. The average connectivity $`<k>`$ is obtained with continuous approximation, $`<k>`$ $`=`$ $`{\displaystyle _m^K}kp(k)𝑑k={\displaystyle _m^K}kck^\alpha 𝑑k={\displaystyle \frac{c}{2\alpha }}(K^{2\alpha }m^{2\alpha })`$ (6) and thus $$<k>\frac{\alpha 1}{2\alpha }m(N^{\frac{2\alpha }{\alpha 1}}1)$$ (7) According to the above results, we can formulate the scale-free network’s resilience to random failures problem as the following nonlinear mixed integer programming. $`max`$ $`H(\alpha ,m,N)`$ $`s.t.`$ $`<k>=const`$ The problem mentioned above is different from the usual maximizing the entropy of the network problem. Note that the highest entropy for a network with a given $`<k>`$ is exponential random graph . To get the optimal design of scale-free networks, we restrict the degree distribution to be a power-law distribution. The problem can be rewritten as follows: $`max`$ $`H(\alpha ,m,N)`$ $`s.t.`$ $`{\displaystyle \frac{\alpha 1}{2\alpha }}m(N^{\frac{2\alpha }{\alpha 1}}1)=const`$ (9) $`2<\alpha <3`$ $`1m(N1)`$ $`minteger`$ By solving the mixed integer programming problem, we get the optimal value $`H`$, the optimal solutions $`\alpha `$ and $`m`$ for different average connectivity $`<k>`$. The results are shown in Fig. 2. From Fig. 2, we can see that, in general, with the increase of average connectivity $`<k>`$, entropy $`H`$ as well as the minimal connectivity $`m`$ increases, too. In part, the optimal minimal connectivity $`m`$ keeps constant for a certain range of $`<k>`$, where scaling exponent $`\alpha `$ decreases with the increase of $`<k>`$. As a result, the optimal network becomes much more heterogeneous than the one with bigger scaling exponent. To prove the entropy of the degree distribution is indeed an effective measure of network robustness under node removal, we examine the relationship between the entropy of the degree distribution and the network threshold value $`p_c`$, which was presented by Cohen *et al.* with percolation theory . The criterion to calculate the percolation critical threshold value $`p_c`$ of a randomly connected network is: $$1p_c=\frac{1}{\kappa _01}$$ (10) where $`\kappa _0=\frac{<k^2>}{<k>}`$ is calculated from the original distribution before the random failures. With $`KmN^{\frac{1}{\alpha 1}}`$, $`\kappa _0`$ can be expressed as follows: $$\kappa _0\frac{\alpha 2}{3\alpha }m(N^{\frac{3\alpha }{\alpha 1}}1)$$ (11) In Fig. 3, $`p_c`$ is calculated from the optimal value $`\alpha `$ and $`m`$ which are obtained from maximizing $`H`$ for different average connectivity $`<k>`$. The results indicate that for some constant $`m`$, $`p_c`$ increases with the increase of $`<k>`$. Since both $`\alpha `$ and $`N`$ affect network’s resilience to random failures, we may ask how the network’s robustness changes with $`N`$. We just examine the relationship between $`p_c`$ and $`\alpha `$ for different $`N`$. The results are shown in Fig. 4. From Fig. 4, we can see that with the increase of $`\alpha `$, $`p_c`$ decreases, while the threshold value $`p_c`$ becomes bigger with the increase of network scale $`N`$, which are consistent with the results shown in Fig.1. To get an intuitive understanding of the relationship between the two measures, we show the results of the threshold value $`p_c`$ and $`H`$ for different average connectivity $`<k>`$ in Fig. 5. It can be seen that $`p_c`$ and $`H`$ vary consistently, that is, $`p_c`$ increases with $`H`$ for different average degree $`<k>`$. Also, with the increase of $`<k>`$, both $`p_c`$ and $`H`$ get improved. ## 3 Conclusions In real-world networks one is often interested in the question of how to design a robust network while keeping the cost constant. It is well known that scale-free networks are resilient to random failures for their heterogeneous links distributions. In this paper, different from percolation theory to evaluate the threshold value $`p_c`$, we use the entropy of the degree distribution to study the resilience of scale-free networks to random failures. Heterogeneity is a simple but essential character of scale-free networks and it is in direct relationship with the network’s resilience to random failures. This character motivates us to use the entropy of the degree distribution, which provides an average measure of heterogeneity to study the effect of random failures on scale-free networks. By optimizing the entropy of the power-law degree distribution, we get the optimal value of entropy $`H`$, the optimal solutions of $`\alpha `$ and $`m`$ for different $`<k>`$. The results indicate that when the network scale $`N`$ is given, the optimal solution $`m`$ is constant for a certain range of $`<k>`$, in which range the scaling exponent $`\alpha `$ decreases. So a higher entropy is obtained. Of course, large average connectivity brings high cost. We then examine the relationship between threshold value $`p_c`$ and $`N`$ for different $`\alpha `$. The results are also consistent with that of the entropy of the degree distribution. We also show the relationship of the two measures$``$$`p_c`$ and $`H`$. The results indicate that $`p_c`$ and $`H`$ vary consistently, that is, $`p_c`$ increases with $`H`$ for a certain average degree $`<k>`$. Also, both $`p_c`$ and $`H`$ get improved with the increase of the average degree $`<k>`$. We just try an alternative point of view to analyze the robustness of the network from its heterogeneity. The results indicate that the entropy of the degree distribution is an effective measure of scale-free network’s resilience to random failures. Deep discussions on the resilience of scale-free networks to targeted attacks will be made in the future. ## Acknowledgements We are grateful to the anonymous referees for their insightful and helpful comments. The work was supported by NSFC(90103033).
warning/0506/astro-ph0506597.html
ar5iv
text
# Hypernovae, Black-Hole-Forming Supernovae, and First Stars ## 1. Introduction and Summary Stars more massive than $``$ 25 $`M_{}`$ form a black hole at the end of their evolution. Stars with non-rotating black holes are likely to collapse ”quietly” ejecting a small amount of heavy elements (Faint supernovae). In contrast, stars with rotating black holes are likely to give rise to very energetic supernovae (Hypernovae). We present distinct nucleosynthesis features of these two classes of ”black-hole-forming” supernovae. Nucleosynthesis in Hypernovae is characterized by larger abundance ratios (Zn,Co,V,Ti)/Fe and smaller (Mn,Cr)/Fe than normal supernovae, which can explain the observed trend of these ratios in extremely metal-poor stars. Nucleosynthesis in Faint supernovae is characterized by a large amount of fall-back. We show that the abundance pattern of the recently discovered most Fe-poor star, HE0107-5240, and other extremely metal-poor stars are in good accord with those of black-hole-forming supernovae, but not pair-instability supernovae. This suggests that black-hole-forming supernovae made important contributions to the early Galactic (and cosmic) chemical evolution. Finally we discuss the nature of First (Pop III) Stars. ## 2. Hypernovae and Faint Supernovae Type Ic Hypernovae 1998bw and 2003dh were clearly linked to the Gamma-Ray Bursts GRB 980425 (Galama et al. 1998) and GRB 030329 (Stanek et al. 2003; Hjorth et al. 2003), thus establishing the connection between long GRBs and core-collapse supernovae (SNe). SNe 1998bw and 2003dh were exceptional for SNe Ic: they were as luminous at peak as a SN Ia, indicating that they synthesized 0.3 - 0.5 $`M_{}`$ of <sup>56</sup>Ni, and their kinetic energy (KE) were estimated as $`E_{51}=E/10^{51}`$ erg $``$ 30 (Iwamoto, Mazzali, Nomoto, et al. 1998; Woosley, Eastman, & Schmidt 1999; Nakamura et al. 2001a; Mazzali et al. 2003). Other “hypernovae” have been recognized, such as SN 1997ef (Iwamoto et al. 2000; Mazzali, Iwamoto, & Nomoto 2000), SN 1999as (Knop et al. 1999; Hatano et al. 2001), and SN 2002ap (Mazzali et al. 2002). These hypernovae span a wide range of properties, although they all appear to be highly energetic compared to normal core-collapse SNe. The mass estimates, obtained from fitting the optical light curves and spectra, place hypernovae at the high-mass end of SN progenitors. In contrast, SNe II 1997D and 1999br were very faint SNe with very low KE (Turatto et al. 1998; Hamuy 2003; Zampieri et al. 2003). In the diagram that shows $`E`$ and the mass of <sup>56</sup>Ni ejected $`M(^{56}`$Ni) as a function of the main-sequence mass $`M_{\mathrm{ms}}`$ of the progenitor star (Figure 1), therefore, we propose that SNe from stars with $`M_{\mathrm{ms}}\text{ }>`$ 20-25 $`M_{}`$ have different $`E`$ and $`M(^{56}`$Ni), with a bright, energetic “hypernova branch” at one extreme and a faint, low-energy SN branch at the other (Nomoto et al. 2003). For the faint SNe, the explosion energy was so small that most <sup>56</sup>Ni fell back onto the compact remnant. Thus the faint SN branch may become a “failed” SN branch at larger $`M_{\mathrm{ms}}`$. Between the two branches, there may be a variety of SNe (Hamuy 2003). This trend might be interpreted as follows. Stars with $`M_{\mathrm{ms}}\text{ }<`$ 20-25 $`M_{}`$ form a neutron star, producing $``$ 0.08 $`\pm `$ 0.03 $`M_{}`$ <sup>56</sup>Ni as in SNe 1993J, 1994I, and 1987A. Stars with $`M_{\mathrm{ms}}\text{ }>`$ 20-25 $`M_{}`$ form a black hole; whether they become hypernovae or faint SNe may depend on the angular momentum in the collapsing core, which in turn depends on the stellar winds, metallicity, magnetic fields, and binarity. Hypernovae might have rapidly rotating cores owing possibly to the spiraling-in of a companion star in a binary system. ## 3. Synthesis of <sup>56</sup>Ni in $`100M_{}`$ Stars The light curve modeling of the unusually bright hypernova SN1999as suggests that the progenitor is a core-collapse supernova and the ejected <sup>56</sup>Ni mass is as large as $`34M_{}`$. Motivated by SN 1990as, Umeda & Nomoto (2004) have investigated how much <sup>56</sup>Ni can be synthesized in core-collapse massive supernovae. The evolutions of several very massive stars with initial masses of $`M100M_{}`$ and low metallicity ($`Z=Z_{}/200`$) have been calculated from the main-sequence to “hypernova” explosions. The synthesized <sup>56</sup>Ni mass increases with the increasing explosion energy and the progenitor mass. Umeda & Nomoto (2004) found that for the explosion energy of 3$`\times 10^{52}`$ ergs, for example, the <sup>56</sup>Ni mass of up to 2.2, 2.3, 5.0, and 6.6 $`M_{}`$ can be produced for the progenitors with masses of 30, 50, 80 and 100 $`M_{}`$, that are sufficiently large to explain SN 1999as. Figure 2 shows the evolution of the central density and temperature for the 30 and 90$`M_{}`$ models. More massive stars have larger specific entropy at the center, thus having higher temperature for the same density. For 90$`M_{}`$, the evolutinary track is very close to (but outside of) the “e<sup>-</sup>e<sup>+</sup> pair-instabillity region” of $`\mathrm{\Gamma }<4/3`$ where $`\mathrm{\Gamma }`$ denotes the adiabatic index. The evolution of the central temperature and density is significantly different between the 30 and 90$`M_{}`$ models during Si-burning at $`T_9=T/10^9`$K =$`2.54`$. The central temperature and density of the 90$`M_{}`$ model oscillate several times. This is because in such massive stars radiation pressure is so dominant that $`\mathrm{\Gamma }`$ is close to 4/3, and thus the inner core of the stars easily expands with the nuclear energy released by Si-burning. Once it expands, the temperature drops suddenly, the central Si-burning stops, and the stellar core turns into shrink. Since only small amount of Si is burnt for each cycle, this pulsations occur many times. Umeda & Nomoto (2004) found from the study of 80$`100M_{}`$ stars that the number of the oscillations depends on the convective parameter $`f_k`$: larger $`f_k`$ increases the number of the oscillation. This is because for a larger $`f_k`$ fresh Si is mixed more efficiently into the center, which increase the lifetime of this stage. The situation is similar to the breathing pulse phase at the end of He-burning, while the variation of the temperature and density is much larger in Si burning than in the breathing pulse phase. The amplitude of the temperature and density variation is larger for more massive stars, which suggests more and more drastic oscillations would occur for larger mass stars. ## 4. Nucleosynthesis in Hypernovae In core-collapse supernovae/hypernovae, stellar material undergoes shock heating and subsequent explosive nucleosynthesis. Iron-peak elements are produced in two distinct regions, which are characterized by the peak temperature, $`T_{\mathrm{peak}}`$, of the shocked material. For $`T_{\mathrm{peak}}>5\times 10^9`$K, material undergoes complete Si burning whose products include Co, Zn, V, and some Cr after radioactive decays. For $`4\times 10^9`$K $`<T_{\mathrm{peak}}<5\times 10^9`$K, incomplete Si burning takes place and its after decay products include Cr and Mn (Nakamura et al. 1999). ### 4.1. Supernovae vs. Hypernovae The right panel of Figure 3 shows the composition in the ejecta of a 25 $`M_{}`$ hypernova model ($`E_{51}=10`$). The nucleosynthesis in a normal 25 $`M_{}`$ SN model ($`E_{51}=1`$) is also shown for comparison in the left panel of Figure 3 (Umeda & Nomoto 2002). We note the following characteristics of nucleosynthesis with very large explosion energies (Nakamura et al. 2001b; Nomoto et al. 2001; Umeda & Nomoto 2005): (1) Both complete and incomplete Si-burning regions shift outward in mass compared with normal supernovae, so that the mass ratio between the complete and incomplete Si-burning regions becomes larger. As a result, higher energy explosions tend to produce larger \[(Zn, Co, V)/Fe\] and smaller \[(Mn, Cr)/Fe\], which can explain the trend observed in very metal-poor stars (Umeda & Nomoto 2005). (2) In the complete Si-burning region of hypernovae, elements produced by $`\alpha `$-rich freezeout are enhanced. Hence, elements synthesized through capturing of $`\alpha `$-particles, such as <sup>44</sup>Ti, <sup>48</sup>Cr, and <sup>64</sup>Ge (decaying into <sup>44</sup>Ca, <sup>48</sup>Ti, and <sup>64</sup>Zn, respectively) are more abundant. (3) Oxygen burning takes place in more extended regions for the larger KE. Then more O, C, Al are burned to produce a larger amount of burning products such as Si, S, and Ar. Therefore, hypernova nucleosynthesis is characterized by large abundance ratios of \[Si,S/O\], which can explain the abundance feature of M82 (Umeda et al. 2002). ### 4.2. Hypernovae and Zn, Co, Mn, Cr Hypernova nucleosynthesis may have made an important contribution to Galactic chemical evolution. In the early galactic epoch when the galaxy was not yet chemically well-mixed, \[Fe/H\] may well be determined by mostly a single SN event (Audouze & Silk 1995). The formation of metal-poor stars is supposed to be driven by a supernova shock, so that \[Fe/H\] is determined by the ejected Fe mass and the amount of circumstellar hydrogen swept-up by the shock wave (Ryan, Norris, & Beers 1996). Then, hypernovae with larger $`E`$ are likely to induce the formation of stars with smaller \[Fe/H\], because the mass of interstellar hydrogen swept up by a hypernova is roughly proportional to $`E`$ (Ryan et al. 1996; Shigeyama & Tsujimoto 1998) and the ratio of the ejected iron mass to $`E`$ is smaller for hypernovae than for normal supernovae. In the observed abundances of halo stars, there are significant differences between the abundance patterns in the iron-peak elements below and above \[Fe/H\]$`2.5`$ \- $`3`$. (1) For \[Fe/H\]$`\text{ }<2.5`$, the mean values of \[Cr/Fe\] and \[Mn/Fe\] decrease toward smaller metallicity, while \[Co/Fe\] increases (McWilliam et al. 1995; Ryan et al. 1996). (2) \[Zn/Fe\]$`0`$ for \[Fe/H\] $`3`$ to $`0`$ (Sneden, Gratton, & Crocker 1991), while at \[Fe/H\] $`<3.3`$, \[Zn/Fe\] increases toward smaller metallicity (Cayrel et al. 2004). The larger \[(Zn, Co)/Fe\] and smaller \[(Mn, Cr)/Fe\] in the supernova ejecta can be realized if the mass ratio between the complete Si burning region and the incomplete Si burning region is larger, or equivalently if deep material from the complete Si-burning region is ejected by mixing or aspherical effects. This can be realized if (1) the mass cut between the ejecta and the compact remnant is located at smaller $`M_r`$ (Nakamura et al. 1999), (2) $`E`$ is larger to move the outer edge of the complete Si burning region to larger $`M_r`$ (Nakamura et al. 2001b), or (3) asphericity in the explosion is larger. Among these possibilities, a large explosion energy $`E`$ enhances $`\alpha `$-rich freezeout, which results in an increase of the local mass fractions of Zn and Co, while Cr and Mn are not enhanced (Umeda & Nomoto 2002, 2005). Models with $`E_{51}=1`$ do not produce sufficiently large \[Zn/Fe\]. To be compatible with the observations of \[Zn/Fe\] $`0.5`$, the explosion energy must be much larger, i.e., $`E_{51}\text{ }>20`$ for $`M\text{ }>20M_{}`$, i.e., hypernova-like explosions of massive stars ($`M\text{ }>25M_{}`$) with $`E_{51}>10`$ are responsible for the production of Zn. In the hypernova models, the overproduction of Ni, as found in the simple “deep” mass-cut model, can be avoided (Umeda & Nomoto 2005). Therefore, if hypernovae made significant contributions to the early Galactic chemical evolution, it could explain the large Zn and Co abundances and the small Mn and Cr abundances observed in very metal-poor stars (Fig. 4: Tominaga et al. 2005). ## 5. Extremely Metal-Poor (EMP) Stars Recently the most Fe deficient and C-rich low mass star, HE0107-5240, was discovered (Christlieb et al. 2002, 2004). This star has \[Fe/H\] $`=5.3`$ but its mass is as low as 0.8 $`M_{}`$. This would challenge the recent theoretical arguments that the formation of low mass stars, which should survive until today, is suppressed below \[Fe/H\] $`=4`$ (Schneider et al. 2002). The important clue to this problem is the observed abundance pattern of this star. This star is characterized by a very large ratios of \[C/Fe\] = 4.0 and \[N/Fe\] = 2.3, while the abundances of elements heavier than Mg are as low as Fe (Christlieb et al. 2002). Interestingly, this is not the only extremely metal poor (EMP) stars that have the large C/Fe and N/Fe ratios, but several other such stars have been discovered (Aoki et al. 2002). Therefore the reasonable explanation of the abundance pattern should explain other EMP stars as well. We show that the abundance pattern of C-rich EMP stars can be reasonably explained by the nucleosynthesis of 20 - 130 $`M_{}`$ supernovae with various explosion energies and the degree of mixing and fallback of the ejecta. ### 5.1. The Most Fe-Poor Star HE0107-5240 We consider a model that C-rich EMP stars are produced in the ejecta of (almost) metal-free supernova mixed with extremely metal-poor interstellar matter. We use Pop III pre-supernova progenitor models, simulate the supernova explosion and calculate detailed nucleosynthesis (Umeda & Nomoto 2003). In Figure 5 (right) we show that the elemental abundances of one of our models are in good agreement with HE0107-5240, where the progenitor mass is 25 $`M_{}`$ and the explosion energy $`E_{51}=`$ 0.3 (Umeda & Nomoto 2003). In this model, explosive nucleosynthesis takes place behind the shock wave that is generated at $`M_r=`$ 1.8 $`M_{}`$ and propagates outward. The resultant abundance distribution is seen in Figure 5 (left), where $`M_r`$ denotes the Lagrangian mass coordinate measured from the center of the pre-supernova model. The processed material is assumed to mix uniformly in the region from $`M_r=`$ 1.8 $`M_{}`$ and 6.0 $`M_{}`$. Such a large scale mixing was found to take place in SN1987A and various explosion models (Hachisu et al. 1990). Almost all materials below $`M_r=`$ 6.0 $`M_{}`$ fall back to the central remnant and only a small fraction ($`f=2\times `$ 10<sup>-5</sup>) is ejected from this region. The ejected Fe mass is 8 $`\times `$ 10<sup>-6</sup> $`M_{}`$. The CNO elements in the ejecta were produced by pre-collapse He shell burning in the He-layer, which contains 0.2 $`M_{}`$ <sup>12</sup>C. Mixing of H into the He shell-burning region produces 4 $`\times `$ 10<sup>-4</sup> $`M_{}`$ <sup>14</sup>N. On the other hand, only a small amount of heavier elements (Mg, Ca, and Fe-peak elements) are ejected and their abundance ratios are the average in the mixed region. The sub-solar ratios of \[Ti/Fe\] $`=0.4`$ and \[Ni/Fe\] $`=0.4`$ are the results of the relatively small explosion energy ($`E_{51}=`$ 0.3). With this ”mixing and fallback”, the large C/Fe and C/Mg ratios observed in HE0107-5240 are well reproduced. In this model, N/Fe appears to be underproduced. However, N can be produced inside the EMP stars through the C-N cycle, and brought up to the surface during the first dredge up stage while becoming a red-giant star (Boothroyd & Sackmann 1999). ### 5.2. EMP Stars from VLT Observations The ”mixing and fall back” process can reproduce the abundance pattern of the typical EMP stars without enhancement of C and N. Figure 6 shows that the averaged abundances of \[Fe/H\] $`=3.7`$ stars in Cayrel et al. (2003) can be fitted well with the hypernova model of 25 $`M_{}`$ and $`E_{51}=`$ 20 (left) but not with the normal SN model of 15 $`M_{}`$ and $`E_{51}=`$ 1 (right) (Tominaga et al. 2005). ## 6. First Stars It is of vital importance to identify the first generation stars in the Universe, i.e., totally metal-free, Pop III stars. The impact of the formation of Pop III stars on the evolution of the Universe depends on their typical masses. ### 6.1. High Mass vs. Low Mass Recent numerical models have shown that, the first stars are as massive as $``$ 100 $`M_{}`$ (Abel et al. 2002). The formation of long-lived low mass Pop III stars may be inefficient because of slow cooling of metal free gas cloud, which is consistent with the failure of attempts to find Pop III stars. If the most Fe deficient star, HE0107-5240, is a Pop III low mass star that has gained its metal from a companion star or interstellar matter (Yoshii 1981), would it mean that the above theoretical arguments are incorrect and that such low mass Pop III stars have not been discovered only because of the difficulty in the observations? Based on the results in the earlier section, we propose that the first generation supernovae were the explosion of $``$ 20-130 $`M_{}`$ stars and some of them produced C-rich, Fe-poor ejecta. Then the low mass star with even \[Fe/H\] $`<5`$ can form from the gases of mixture of such a supernova ejecta and the (almost) metal-free interstellar matter, because the gases can be efficiently cooled by enhanced C and O (\[C/H\] $`1`$). ### 6.2. Pair Instability SNe vs. Core Collapse SNe In contrast to the core-collapse supernovae of 20-130 $`M_{}`$ stars, the observed abundance patterns cannot be explained by the explosions of more massive, 130 - 300 $`M_{}`$ stars. These stars undergo pair-instability supernovae (PISNe) and are disrupted completely (e.g., Umeda & Nomoto 2002; Heger & Woosley 2002), which cannot be consistent with the large C/Fe observed in HE0107-5240 and other C-rich EMP stars. The abundance ratios of iron-peak elements (\[Zn/Fe\] $`<0.8`$ and \[Co/Fe\] $`<0.2`$) in the PISN ejecta (Fig. 6; Umeda & Nomoto 2002; Heger & Woosley 2002) cannot explain the large Zn/Fe and Co/Fe in the typical EMP stars (McWilliam et al. 1995; Norris et al. 2001; Cayrel et al. 2003) and CS22949-037 either. Therefore the supernova progenitors that are responsible for the formation of EMP stars are most likely in the range of $`M20130`$ $`M_{}`$, but not more massive than 130 $`M_{}`$. This upper limit depends on the stability of massive stars as will be discussed below. ### 6.3. Stability and Mass Loss of Massive Pop III Stars To determine the upper limit mass of the Zero Age Main Sequence (ZAMS), Nomoto et al. (2003) analyzed a linear non-adiabatic stability of massive ($`80M_{}`$ \- $`300M_{}`$) Pop III stars using a radial pulsation code. Because CNO elements are absent during the early stage of their evolution, the CNO cycle does not operate and the star contracts until temperature rises sufficiently high for the $`3\alpha `$ reaction to produce <sup>12</sup>C. We calculate that these stars have $`X_{\mathrm{CNO}}1.64.0\times 10^{10}`$, and the central temperature $`T_c1.4\times 10^8K`$ on their ZAMS. We also examine the models of Pop I stars for comparison. Table 1 shows the results for our analysis. The critical mass of ZAMS Pop III star is $`128M_{}`$ while that of Pop I star is $`94M_{}`$. This difference comes from very compact structures (with high $`T_c`$) of Pop III stars. Stars more massive than the critical mass will undergo pulsation and mass loss. We note that the $`e`$-folding time of instability is much longer for Pop III stars than Pop I stars with the same mass, and thus the mass loss rate is much lower. These results are consistent with Ibrahim et al. (1981) and Baraffe et al. (2001). However, the absence of the indication of PISNe in EMP stars might imply that these massive stars above 130$`M_{}`$ undergo significant mass loss, thus evolving into Fe core-collapse rather than PISNe. ## 7. Type Ia/IIn Supernovae: SNe 2002ic, 1997cy, and 1999E SNe 1997cy and 1999E were initially classified as Type IIn because they showed H$`\alpha `$ emission. SN 2002ic would also have been so classified, had it not been discovered at an early epoch. SN 1997cy ($`z=0.063`$) is among the most luminous SNe discovered so far ($`M_V<20.1`$ about maximum light), and SN 1999E is also bright ($`M_V<19.5`$). Both SNe 1997cy and 1999E have been suspected to be spatially and temporally related to a GRB (Germany et al. 2000; Rigon et al. 2003). However, both the classification and the associations with a GRB must now be seen as highly questionable in view of the fact that their replica, SN 2002ic, appears to have been a genuine SN Ia at an earlier phase (Hamuy et al. 2003; Deng et al. 2004). Nomoto et al. (2005) calculated the interaction between the expanding ejecta and CSM (circumstellar matter). For the supernova ejecta, the carbon deflagration model W7 was used. After day $``$ 350, the light curve starts declining. To reproduce the declining part of the light curve, we add the outer CSM of 0.2 $`M_{}`$ where the density declines sharply as $`n=6`$. This implies that the total mass of CSM is $``$ 1.3 $`M_{}`$.
warning/0506/astro-ph0506056.html
ar5iv
text
# Primordial non-gaussianities from multiple-field inflation ## 1 Introduction The inflationary paradigm is an attractive proposal for the evolution of the very early universe . During inflation the universe underwent a phase of accelerated expansion driven by one or more self-interacting scalar fields. Despite the appealing simplicity behind the central idea of inflation, it has proved difficult to discriminate between the large number of different models that have been developed to date . The simplest classes of models typically predict that the spectrum of primordial density perturbations generated quantum-mechanically during inflation should be nearly scale-invariant and Gaussian-distributed. (For a review, see, e.g., Ref. .) These generic predictions are in good agreement with cosmological observations, most notably those arising from measurements of the temperature anisotropy and polarization of the Cosmic Microwave Background (CMB) . Such observations lead to strong constraints on theoretical model building. However, if further progress is to be made in our understanding of the microphysics of the early universe, it will become increasingly necessary to extend the theoretical framework beyond the leading-order effects of scale-invariant, Gaussian fluctuations. In particular, deviations away from Gaussian statistics represent a potentially powerful discriminant between competing inflationary models and have attracted considerable recent interest . Moreover, measurements of CMB anisotropies are now approaching a level of precision where such differences could soon be detectable and, indeed, the first non-gaussian signals in the data may already have been observed (see also ). In general, any Gaussian random variable $`X`$ with zero mean and variance $`\sigma ^2`$ has a probability measure given by $$\left[X(x,x+\mathrm{d}x)\right]\mathrm{d}x\mathrm{exp}\left(\frac{x^2}{2\sigma ^2}\right)\mathrm{d}x,$$ (1) and the correlation functions of $`X`$ have the form $$X^n=x^n\left[X(x,x+\mathrm{d}x)\right]dx.$$ (2) The rules of Gaussian integration imply that $`X^n`$ vanishes whenever $`n`$ is odd, whereas $`X^n`$ can be written in terms of $`X^2`$ for even $`n`$. Accordingly, we expect violations of Gaussian statistics to manifest themselves as deviations from these simple rules. The same principle holds when dealing with fluctuations $`\delta \phi `$ in some quantum field, where the semiclassical probability density is given by $`(\delta \phi )\mathrm{e}^{I[\delta \phi ]}`$ and $`I`$ is the Euclidean action. When $`I`$ describes a free theory with no interactions, $`\delta \phi `$ may be viewed as a Gaussian random variable. However, if interactions are present, deviations from pure Gaussian statistics will arise. More specifically, if $`\delta \phi `$ represents an ‘almost Gaussian’ random variable, then it is expected (in a largely model-independent sense) that the three-point function $`\delta \phi ^3`$ will provide the largest contribution to the non-gaussianity. In typical single- or multi-field inflationary models, the level of non-gaussianity can be calculated analytically . The single-field scenario has been well-studied , and some effort has focused recently on non-standard theories of gravity, in the hope of identifying observational signatures that differ from those of canonical inflation based on the Einstein action . Some earlier investigations considered non-gaussian fluctuations, but were specifically directed at the formation of primordial black holes during the final stages of inflation, when the slow-roll approximation has broken down . In general, however, the superhorizon evolution of the curvature perturbation during single-field inflation is simple, and there is a robust prediction that the level of primordial non-gaussianity as measured by the three-point correlation function cannot be large , even allowing for exotic effects such as higher-derivative operators of the scalar field or a speed of sound that is different from that of light . After the end of inflation, one must evolve this small non-gaussianity through the epochs of radiation and matter domination in order to arrive at the temperature anisotropy $`\delta T/T`$ which is seen on the microwave background sky . This evolution after the end of inflation gives an additional source of non-gaussianity which is comparable to the primordial non-gaussianity in the case of single-field inflation. By comparison, our understanding of the multi-field scenario is less developed, although some progress has been made . Typically, the evolution of the curvature perturbation on superhorizon scales implies that a significant non-gaussian signal can in principle be generated during inflation. Indeed, a fairly general prescription for calculating the three-point function of fluctuations was recently presented , in which it was assumed that the curvature perturbation evolves substantially outside the horizon due to pressure effects after the end of inflation. Such superhorizon evolution effects generically arise when more than one fluid is cosmologically important, or when a single fluid is present but with an indefinite equation of state. In the multiple-field case, therefore, the primordial three-point function consists of a microphysical component, calculated from the properties of the matter fluctuations around the epoch of horizon exit, and a superhorizon component, which is determined purely by large-scale gravitational effects. If the superhorizon component becomes large, which typically will not happen until after the end of inflation, it may dominate the primordial non-gaussianity, in which case it is consistent to neglect the microphysical contribution. This actually occurs, for example, in the curvaton scenario . In this respect, the theory has some similarities with the stochastic approach to inflation, which has been revived recently within the context of non-gaussianity . However, in the absence of any estimate for the microphysical contribution to the three-point function, the above prescription is necessarily incomplete, since it is not possible to verify explicitly that the superhorizon contribution dominates. In this paper, we supply the missing ingredient by calculating the three-point function evaluated at the epoch of horizon crossing for a generic set of $`𝒩`$ scalar fields $`\{\phi ^I\}`$ coupled to gravity. We consider a (Lorentzian) action of the form $$S=\frac{1}{2}\mathrm{d}^4x\sqrt{g}\left[R𝒢_{IJ}_\mu \phi ^I^\mu \phi ^J2V(\phi )\right].$$ (3) where the interaction potential $`V(\phi )`$ is assumed to be an arbitrary function of the fields, $`\phi ^I`$, and $`𝒢_{IJ}`$ represents the metric on the manifold parametrized by the scalar field values. (We refer to this metric as the ‘target space’ metric). The action for $`𝒩`$ canonically normalized scalar fields minimally coupled to gravity is recovered by specifying $`𝒢_{IJ}=\delta _{IJ}`$, where $`\delta _{IJ}`$ represents the Kronecker-delta. More generally, the kinetic sector of (3) may be invariant under a global symmetry group $`J`$, in which case the scalar fields parametrize the coset space $`J/K`$ with a line element given by $`\mathrm{d}s^2=G_{IJ}\mathrm{d}\phi ^I\mathrm{d}\phi ^J`$, where $`K`$ is the maximal compact subgroup of $`J`$. In such models, the metric components generically exhibit a direct dependence on the scalar fields $`\phi ^I`$. In this paper, we restrict our attention to the case where the target space is approximately flat, at least over the region dynamically explored by the fields $`\phi ^I`$, and the metric can be brought to the field-*independent* form $`𝒢_{IJ}=\delta _{IJ}`$ by an appropriate choice of parametrization. On the other hand, we allow the potential $`V(\phi )`$ to be quite arbitrary, subject to the usual slow-roll conditions. This choice of action covers a wide class of multi-field inflationary scenarios. The outline of the paper is as follows. In Section 2, we briefly describe how to calculate the primordial non-gaussianity produced during multi-field inflation, as measured by the three-point function of the curvature perturbation. The crucial ingredients for the calculation are: (i) an expression for the curvature perturbation itself, which is provided by the Sasaki–Stewart $`\delta N`$ formalism , and: (ii) an estimate of the scalar three-point function just after horizon exit. This three-point function is derived in Section 3 and is presented in Eqs. (68)–(69). It represents the central result of the paper. In Section 4, we verify that our formalism reproduces the well-known result of Maldacena when specialised to the case of a single scalar field. In Section 5 the non-gaussianity generated during assisted inflation is calculated as an example of our method. Finally, we conclude in Section 6. Units are chosen such that $`c=\mathrm{}=M_\mathrm{P}=1`$, where $`M_\mathrm{P}=(8\pi G)^{1/2}`$ represents the reduced Planck mass. ## 2 Non-gaussianity from multiple fields ### 2.1 The background model We assume throughout that the background model corresponds to the spatially flat Friedmann–Robertson–Walker (FRW) universe, with a metric $$\mathrm{d}s^2=\mathrm{d}t^2+a^2\delta _{ij}\mathrm{d}x^i\mathrm{d}x^j.$$ (4) The scalar field equations take the form of a set of coupled Klein–Gordon equations : $$\ddot{\phi }^I+3H\dot{\phi }^I+𝒢^{IJ}V_{,J}=0,$$ (5) where $`H=\dot{a}/a`$ is the Hubble parameter, $`a`$ is the scale factor, $`\dot{\phi }^I=\mathrm{d}\phi ^I/\mathrm{d}t`$ and a comma denotes partial differentiation. The gravitational equation of motion is $$2\dot{H}+3H^2=\frac{1}{2}𝒢_{IJ}\dot{\phi }^I\dot{\phi }^J+V(\phi ),$$ (6) together with the Friedmann constraint equation $$3H^2=\frac{1}{2}𝒢_{IJ}\dot{\phi }^I\dot{\phi }^J+V(\phi ).$$ (7) Eqs. (6) and (7) together imply that $$\dot{H}=\frac{1}{2}𝒢_{IJ}\dot{\phi }^I\dot{\phi }^J.$$ (8) There is also an analogue of the single-field Hamilton–Jacobi equation, which reads $$H_{,I}=\frac{1}{2}\dot{\phi }_I.$$ (9) As in single-field inflation, it will prove convenient to introduce a set of slow-roll parameters. One such parameter is the slow-roll matrix $`\epsilon ^{IJ}`$: $$\epsilon ^{IJ}=\frac{\dot{\phi }^I\dot{\phi }^J}{2H^2}=2𝒢^{IM}𝒢^{JN}\frac{H_{,M}H_{,N}}{H^2}=\epsilon ^I\epsilon ^J,$$ (10) where we have used (9), and $`\epsilon ^I`$ is the vector $$\epsilon ^I=\frac{\dot{\phi }^I}{\sqrt{2}H}=\frac{\sqrt{2}𝒢^{IJ}H_{,J}}{H}.$$ (11) The standard slow-roll parameter, $`\epsilon =\dot{H}/H^2`$, can then be expressed in terms of $`\epsilon ^{IJ}`$, since $`\epsilon =\mathrm{tr}\epsilon ^{IJ}=𝒢_{IJ}\epsilon ^{IJ}`$. As a result, we generally expect $`|\epsilon ^{IJ}|1`$ whenever slow-roll is valid except in models where finely-tuned cancellations between the components of $`\epsilon ^{IJ}`$ may occur. Specifically, for generic theories which are not tuned to have anomalous parameter values, one should expect $$\epsilon ^{IJ}=\mathrm{O}\left(\frac{\epsilon }{𝒩}\right),\epsilon ^I=\mathrm{O}\left(\frac{\epsilon ^{1/2}}{𝒩^{1/2}}\right).$$ (12) The time derivative of $`\epsilon ^{IJ}`$ is determined by $$\dot{\epsilon }^{IJ}=2\epsilon H\left(\epsilon ^{IJ}\eta ^{IJ}\right),$$ (13) where we have defined a second slow-roll matrix $$\eta ^{IJ}=\frac{\ddot{\phi }^I\dot{\phi }^J+\dot{\phi }^J\ddot{\phi }^J}{4H\dot{H}}.$$ (14) This matrix generalizes the slow-roll parameter $`\eta `$ of single-field inflation, as can be deduced by substituting in Eq. (8) to yield $`\eta ^{\phi \phi }=\ddot{\phi }/H\dot{\phi }=\eta `$. Finally, we introduce a third matrix defined by $$\stackrel{~}{\eta }_{IJ}=\frac{V_{,IJ}}{3H^2}.$$ (15) This is related to $`\epsilon ^{IJ}`$ and $`\eta ^{IJ}`$ by the simple rule $$\epsilon ^{IJ}+\eta ^{IJ}=\stackrel{~}{\eta }_{MN}\frac{𝒢^{M(I}\epsilon ^{J)N}}{\epsilon }.$$ (16) The parameter $`\stackrel{~}{\eta }_{IJ}`$ represents a generalization of the slow-roll parameter $`V^{\prime \prime }/V`$ of single-field inflation, and can be related to the alternative definition $`\ddot{\phi }/H\dot{\phi }`$. In multi-field models, however, the index arrangement in (16) means that it is convenient to maintain two separate definitions. We note that Eq. (16) implies that $`\eta ^{IJ}`$ and $`\stackrel{~}{\eta }_{IJ}`$ are of the same order in slow-roll. The utility of this formulation of the slow-roll approximation is essentially restricted to the case of a flat target space. It will also be useful to have on hand an explicit expression for the number of e-foldings in this model. One defines the number of e-folds $`N`$ which elapse between some initial value $`a_i`$ of the scale factor and a final value $`a_f`$ as $`N=\mathrm{ln}a_f/a_i`$, so $$\mathrm{d}N=H\mathrm{d}t=\frac{1}{\epsilon }\mathrm{d}\mathrm{ln}H=\frac{\epsilon _I}{\sqrt{2}\epsilon }\mathrm{d}\phi ^I.$$ (17) ### 2.2 The uniform density curvature perturbation We now consider perturbation theory around the above homogeneous background. Since it is assumed that the scalar fields $`\phi ^I`$ dominate the energy density of the universe, any perturbations $`\delta \phi ^I`$ in these fields necessarily produce a disturbance in the energy–momentum tensor, which back-reacts on the spacetime curvature. Therefore, when working with the perturbation theory of these scalar fields, we must also take into account the effect of scalar metric fluctuations in order to include the effect of this back-reaction. To first-order in perturbation theory, these metric fluctuations can be written as $$\mathrm{d}s^2=(1+2\mathrm{\Phi })\mathrm{d}t^2+2a^2B_{,i}\mathrm{d}x^i\mathrm{d}t+a^2\left[(12\mathrm{\Psi })\delta _{ij}+2E_{,ij}\right]\mathrm{d}x^i\mathrm{d}x^j.$$ (18) Not all of these degrees of freedom are dynamically important, however, since they are related by the gauge transformations $`tt+\delta t`$ and $`x^ix^i+\delta x^i`$, which reshuffle $`\mathrm{\Phi }`$, $`\mathrm{\Psi }`$, $`B_{,i}`$ and $`E_{,ij}`$. To obtain a gauge-invariant measure of the strength of such fluctuations, it is sufficient to consider scalar curvature invariants of the three-dimensional hypersurfaces $`t=\text{constant}`$. The simplest such invariant is given by $$\mathrm{\Psi }=\frac{a^2}{4}^2R^{(3)},$$ (19) where $`R^{(3)}`$ is the three-dimensional Ricci scalar on the spatial hypersurfaces, and $`^2`$ is the inverse Laplacian<sup>1</sup><sup>1</sup>1This is most simply defined in Fourier space, where the Laplacian $`^2`$ is represented as multiplication by $`k^2`$. Similarly, the inverse Laplacian corresponds to multiplication by $`k^2`$.. In single-field models, this quantity has an attractive interpretation. Under a temporal gauge transformation $`tt+H\delta t`$, the field $`\mathrm{\Psi }`$ transforms according to the rule $$\mathrm{\Psi }\mathrm{\Psi }+H\delta t.$$ (20) Hence, the curvature scalar on comoving hypersurfaces (defined by the condition that the scalar field perturbation $`\delta \phi `$ vanishes) is given by $$=\mathrm{\Psi }_{\delta \phi =0}=\mathrm{\Psi }+H\frac{\delta \phi }{\dot{\phi }}.$$ (21) One may also consider the curvature scalar on uniform density hypersurfaces, defined by $`\delta \rho =0`$. This scalar is denoted by $`\zeta `$, where $`|\zeta |=|\mathrm{\Psi }_{\delta \rho =0}|`$ , but there is no consistent agreement on sign conventions. For adiabatic perturbations generated during single-field inflation, the two sets of hypersurfaces coincide , so modulo choices for signs, it follows that $`=|\zeta |`$. One can freely work in either the comoving or uniform density gauges, and the gauge-invariant variable $`=|\zeta |=\mathrm{\Psi }+H\delta \phi /\dot{\phi }`$ is the only degree of freedom in the theory. This variable mixes the fluctuations arising from the metric and the scalar field, neither of which has a separate gauge-invariant interpretation. Consequently, at the quantum level one considers the creation of $``$\- or $`\zeta `$-particles from the vacuum rather than that of inflaton particles or gravitons. After horizon exit, $``$ asymptotes towards a time-independent quantity and may be viewed as a function only of position. It can therefore be absorbed by an appropriate choice of coordinates on the spatial hypersurfaces. This implies that distant regions of the universe differ from each other only by how much one has expanded relative to the other. Since this is a symmetry of the equations of motion, $``$ is conserved on superhorizon scales . In multi-field models, the situation is not so simple. It is still convenient to work with the uniform density curvature perturbation, $`\zeta `$, which on large scales is still equivalent to the comoving curvature perturbation $``$. However, expressing $`\zeta `$ as a mixture of the metric perturbation $`\mathrm{\Psi }`$ and the scalar field perturbations $`\delta \phi ^I`$ is much more involved. This was first achieved at a linear level by Sasaki & Stewart , who used (20) to express the difference between an initial uniform curvature hypersurface and a final comoving hypersurface as $$\zeta =|H\mathrm{d}t|=\mathrm{d}N,$$ (22) where $`N`$ is the integrated number of e-folds, as defined in (17). On sufficiently large scales we expect (given suitable assumptions about the dynamical behaviour which permit us to ignore time derivatives of the perturbations) that each horizon volume will evolve as if it were a self-contained universe, and it therefore follows that (as shown beyond linear order by Lyth, Malik & Sasaki ) $$\zeta =N_{,I}\delta \phi ^I+\frac{1}{2}N_{,IJ}\delta \phi ^I\delta \phi ^J+\mathrm{},$$ (23) where the $`\delta \phi ^I`$ express the deviations of the fields from their unperturbed values in some given region of the universe, and we employ the standard summation convention for the scalar field indices $`I`$, …, $`J`$. It is this expression for $`\zeta `$ which was used by Lyth & Rodriguez to express the primordial non-gaussianity of $`\zeta `$ in multi-field models . (See also .) When more than one field is dynamically important, Eq. (23) allows $``$ to evolve outside the horizon while inflation is still occurring<sup>2</sup><sup>2</sup>2To be clear, we should point out that there is evolution of the metric perturbations after the end of inflation. One source of evolution, arising from the evolution of perturbations during the epochs of radiation and matter domination which follow reheating , is universal to all models of inflation, irrespective of whether many fields are important or only one. Additionally, in multiple field models, $`N`$ will typically continue to vary after the end of inflation. Since (23) is still valid, this means that $`\zeta `$ will continue to change. In this paper, we only evaluate $`\zeta `$ up to the end of inflation., in contrast with the single-field case. It can be shown that the evolution of $``$ is sourced by a non-adiabatic pressure perturbation : $$\dot{}\frac{H}{p+\rho }\delta p_{\text{nad}},$$ (24) where $`\delta p_{\text{nad}}`$ can be written in the form $$\delta p_{\text{nad}}=\dot{p}\mathrm{\Gamma }=\dot{p}\left(\frac{\delta p}{\dot{p}}\frac{\delta \rho }{\dot{\rho }}\right),$$ (25) and $`\mathrm{\Gamma }`$ represents the so-called isocurvature perturbation that measures the displacement between hypersurfaces of constant density and of constant pressure. (Note that $`\mathrm{\Gamma }`$ vanishes whenever $`p`$ is a definite function of $`\rho `$.) Thus, whenever isocurvature perturbations are present, which are essentially inevitable in a general multi-field scenario, $``$ will evolve outside the horizon while inflation is still occuring, and this evolution must be taken into account in the inflationary prediction . The Sasaki–Stewart formalism conveniently handles these complexities for us. ### 2.3 The three-point function in multiple-field models In the remainder of this section, we briefly review the results of . In order to connect with observations, any prediction for the non-gaussianity measured by the three-point function must be expressed in terms of an experimentally relevant parameter. A common choice is the non-linearity parameter, $`f_{\mathrm{NL}}`$, which expresses the departure of $`\zeta `$ from a Gaussian random variable: $$\zeta =\zeta _g\frac{3}{5}f_{\mathrm{NL}}\zeta _g^2,$$ (26) where $`\zeta _g`$ is Gaussian and we have written a convolution product since $`f_{\mathrm{NL}}`$ is typically momentum dependent (and defined by multiplication in Fourier space ). We have dropped an uninteresting zero mode in (26), which only serves to fix $`\zeta =0`$. This so-called primordial non-linearity is processed into an observable non-linearity of temperature anisotropies $`f_{\mathrm{NL}}^{\delta T/T}`$ as described in . To relate $`f_{\mathrm{NL}}`$ to the three-point function, we follow Maldacena and write the latter as $$\zeta (𝐤_1)\zeta (𝐤_2)\zeta (𝐤_3)=(2\pi )^3\delta (𝐤_1+𝐤_2+𝐤_3)\frac{4\pi ^4}{_ik_i^3}|\mathrm{\Delta }_\zeta ^2|^2𝒜,$$ (27) where $`𝒜`$ is a momentum-dependent function that can be calculated explicitly, and $`\mathrm{\Delta }_\zeta ^2`$ is the dimensionless power spectrum of $`\zeta `$, defined in the standard way by $$\zeta (𝐤_1)\zeta (𝐤_2)=(2\pi )^3\delta (𝐤_1+𝐤_2)P_\zeta (k_1),\mathrm{\Delta }_\zeta ^2(k)=\frac{k^3}{2\pi ^2}P_\zeta (k),$$ (28) such that $`\zeta \zeta =_0^{\mathrm{}}\mathrm{\Delta }_\zeta ^2(k)\mathrm{d}\mathrm{ln}k`$. The non-linearity parameter can now be written in the succinct form <sup>3</sup><sup>3</sup>3There are other relations between $`f_{\mathrm{NL}}`$ and the three-point function which are common in the literature, usually written in terms of the bispectrum, but they are all equivalent to this one. In the present case, the use of the dimensionless power spectrum is especially convenient because it will allow us to factorize the index structure in the multiple-field three-point functions.: $$f_{\mathrm{NL}}=\frac{5}{6}\frac{𝒜}{_ik_i^3}.$$ (29) It only remains to use (23) to relate $`f_{\mathrm{NL}}`$ and the scalar field three-point functions. We show in Section 3 that the two-point function for the scalars satisfies $$\delta \phi ^I(𝐤_1)\delta \phi ^J(𝐤_2)=(2\pi )^3\delta (𝐤_1+𝐤_2)\frac{2\pi ^2}{k_1^3}\mathrm{\Delta }_{}^2𝒢^{IJ},$$ (30) and that the three-point function can be written in the form $$\delta \phi ^I(𝐤_1)\delta \phi ^J(𝐤_2)\delta \phi ^K(𝐤_3)=(2\pi )^3\delta (𝐤_1+𝐤_2+𝐤_3\frac{4\pi ^4}{_ik_i^3}|\mathrm{\Delta }_{}^2|^2𝒜^{IJK},$$ (31) where $`\mathrm{\Delta }_{}^2`$ is the spectrum of a massless scalar field in de Sitter space and $`𝒜^{IJK}`$ is a momentum-dependent function given in Eq. (69). The determination of $`𝒜^{IJK}`$ represents the principal result of this paper. Eq. (23) implies that the power spectrum for $`\zeta `$ is given by $$\mathrm{\Delta }_\zeta ^2=\mathrm{\Delta }_{}^2N_{,I}N_{,J}𝒢^{IJ}.$$ (32) On the other hand, the three-point function which follows from (23) can be written as $`\zeta (𝐤_1)\zeta (𝐤_2)\zeta (𝐤_3)=N_{,I}N_{,J}N_{,K}\delta \phi ^I(𝐤_1)\delta \phi ^J(𝐤_2)\delta \phi ^K(𝐤_3)`$ $`+{\displaystyle \frac{1}{2}}N_{,I}N_{,K}N_{,MN}\delta \phi ^I(𝐤_1)\delta \phi ^K(𝐤_2)[\delta \phi ^M\delta \phi ^N](𝐤_3)+\text{perms}+\mathrm{},`$ (33) where $``$ denotes the convolution product, and for brevity we have omitted the remaining cross terms in the expansion of $`\zeta ^3`$. (In particular, there will be terms involving the product of a spectrum and a bispectrum, of the form $$\frac{1}{4}N_{,I}N_{,KL}N_{,MN}\delta \phi ^I(𝐤_1)[\delta \phi ^K\delta \phi ^L](𝐤_2)[\delta \phi ^M\delta \phi ^N](𝐤_3)+\text{perms},$$ (34) and a term involving the product of the spectra, of the form $$\frac{1}{8}N_{,IJ}N_{,KL}N_{,MN}[\delta \phi ^I\delta \phi ^J](𝐤_1)[\delta \phi ^K\delta \phi ^L](𝐤_2)[\delta \phi ^M\delta \phi ^N](𝐤_3).$$ (35) The first of these was ignored by Lyth & Rodriguez , because it is proportional to the three-point function for the $`\delta \phi ^I`$, which these authors assumed to vanish. These terms can all be treated similarly when included in $`f_{\mathrm{NL}}`$, but some of the intermediate expressions can become rather long. For these reason we suppress these higher contributions.) The first set of terms in Eq. (23) is just the connected scalar three-point function, since we have assumed that all tadpole contributions vanish. This is mandatory, and equivalent to perturbative stability of the vacuum. Under the same assumptions, the second group of terms can be reduced to products of two-point functions. Consequently, by employing (30)–(31), Eq. (23) can be expressed in the form $$\zeta (𝐤_1)\zeta (𝐤_2)\zeta (𝐤_3)=(2\pi )^3\delta (𝐤_1+𝐤_2+𝐤_3)\frac{4\pi ^4}{_ik_i^3}|\mathrm{\Delta }_{}^2|^2𝒜_\zeta ,$$ (36) where $`𝒜_\zeta `$ is defined by $$𝒜_\zeta =𝒜^{IJK}N_{,I}N_{,J}N_{,K}+𝒢^{IM}𝒢^{KN}N_{,I}N_{,K}N_{,MN}\underset{i}{}k_i^3.$$ (37) Our expression for $`f_{\mathrm{NL}}`$, Eq. (29), now implies that the non-linearity in this model, to a good approximation, is given by $$f_{\mathrm{NL}}=\frac{5}{6}\frac{𝒜^{IJK}N_{,I}N_{,J}N_{,K}}{(N_{,I}N_{,J}𝒢^{IJ})^2_ik_i^3}\frac{5}{6}\frac{𝒢^{IM}𝒢^{KN}N_{,I}N_{,K}N_{,MN}}{(N_{,I}N_{,J}𝒢^{IJ})^2}+\mathrm{},$$ (38) where by ‘$`\mathrm{}`$’ we denote the remaining cross terms from (23) which we did not display. In principle there will also be other terms coming from the Taylor expansion of $`\zeta `$ (in Eq. (23)) beyond second order, but we assume that the series converges sufficiently fast that the effect from these higher-order terms is small. Nevertheless it should be borne in mind that they are present in principle, and in certain models their contribution could be significant. In order to express $`f_{\mathrm{NL}}`$ as in (38), we have exchanged $`\mathrm{\Delta }_{}^2`$ for $`\mathrm{\Delta }_\zeta ^2`$ in (36); this has the effect of introducing the terms involving $`N_{,I}N_{,J}𝒢^{IJ}`$ in the denominators of Eq. (38). We note that the elements of this expression were first given in . Generally speaking, we expect both displayed terms in (38) to be of roughly comparable importance at the end of inflation, although examples of models where one term is dominant do exist. For example, assisted inflation represents a model where the second term vanishes, so $`f_{\mathrm{NL}}`$ arises entirely from the first term. On the other hand, the authors of Ref. focused on the case where the second term is dominant. Note that the first term and any higher cross-terms containing the bispectrum $`𝒜^{IJK}`$ are strongly momentum-dependent, whereas the second term and any higher cross-terms which do not include the bispectrum are momentum-independent, up to mild scale-dependences which arise from renormalization effects . We proceed in the following section to calculate the three-point function for the multi-field action (3). ## 3 The scalar three-point function ### 3.1 Perturbations in the uniform curvature gauge The first step in determining perturbations in multi-field inflation is to select a gauge. The discussion of the previous section, where (22) was used to express the curvature perturbation $`\zeta `$, implies that the relevant gauge should be the uniform curvature slicing, where coordinate time $`t`$ is chosen in such a way that slices of constant $`t`$ have zero Ricci curvature. Such a metric can be written in the ADM form : $$\mathrm{d}s^2=N^2\mathrm{d}t^2+h_{ij}(\mathrm{d}x^i+N^i\mathrm{d}t)(\mathrm{d}x^j+N^j\mathrm{d}t).$$ (39) This version of the metric incorporates the scalar perturbations $`\mathrm{\Phi }`$ and $`B_{,i}`$ written down in (18) within the lapse, $`N`$, and shift, $`N^i`$. These latter two quantities appear as Lagrange multipliers in the action and for this reason it is considerably easier to perform calculations with (39) than with the equivalent expression (18). The $`E_{,ij}`$ term of (18) has been removed by a coordinate redefinition. One specifies $`h_{ij}=a^2(t)\delta _{ij}`$ in order to obtain spatially flat slices. The scalar fields on these flat hypersurfaces are then written as $`\phi ^I=\phi _0^I+Q^I`$, where $`\phi _0^I`$ are the spatially-homogeneous background values of the fields and $`Q^I`$ represent small perturbations. In these coordinates the action (3) becomes $$S=\frac{1}{2}N\sqrt{h}\left(𝒢_{IJ}h^{ij}_i\phi ^I_j\phi ^J2V(\phi )\right)+\frac{1}{2}N^1\sqrt{h}\left(E_{ij}E^{ij}E^2+𝒢_{IJ}v^Iv^J\right),$$ (40) where $`v^I=\dot{\phi }^IN^j_j\phi ^I`$. There are two constraint equations arising from this action. The first follows by demanding that the action be stationary under variations of $`N`$: $$𝒢_{IJ}h^{ij}_i\phi ^I_j\phi ^J2V\frac{1}{N^2}\left(E_{ij}E^{ij}E^2+𝒢_{IJ}v^Iv^J\right)=0.$$ (41) The second constraint arises from the variation in $`N^i`$ and yields $$\left(\frac{1}{N}[E_i^jE\delta _i^j]\right)_{|j}=\frac{1}{N}𝒢_{IJ}v^I_i\phi ^J,$$ (42) where $`|`$ denotes the covariant derivative compatible with the spatial metric $`h_{ij}`$. These constraints can be solved to first order by specifying $`N=1+{\displaystyle \frac{1}{2H}}𝒢_{IJ}\dot{\phi }_0^IQ^J,`$ $`N_i=^2\psi ={\displaystyle \frac{a^2}{2H}}^2𝒢_{IJ}\left(Q^I\ddot{\phi }_0^J\dot{\phi }_0^I\dot{Q}^J{\displaystyle \frac{\dot{H}}{H}}\dot{\phi }_0^IQ^J\right).`$ (43) It turns out that the higher-order pieces in $`N`$ and $`N^i`$ are not required, since they cancel out of the second- and third-order terms in the action (40) . In what follows, we will have no need to refer to the total field $`\phi ^I=\phi _0^I+Q^I`$, and for notational convenience we therefore drop the subscript ‘0’ on $`\phi _0^I`$ and simply identify $`\phi ^I`$ as the homogeneous background field. ### 3.2 The second-order theory After integrating by parts in the action (40) and employing the background field equations (5)–(8) to simplify the result, we obtain the second-order action : $$S_2=\frac{1}{2}dt\mathrm{d}^3xa^3\left(𝒢_{IJ}\dot{Q}^I\dot{Q}^J\frac{1}{a^2}𝒢_{IJ}Q^IQ^J_{IJ}Q^IQ^J\right),$$ (44) where a shorthand notation $`Q^IQ^J`$ is adopted for the scalar product $`\delta ^{ij}_iQ^I_jQ^J`$, and the mass-matrix, $`_{IJ}`$, is defined by $$_{IJ}=V_{,IJ}\frac{1}{a^3}\frac{\mathrm{d}}{\mathrm{d}t}\left(\frac{a^3}{H}\dot{\phi }_I\dot{\phi }_J\right).$$ (45) Eq. (44) is exact and is valid for arbitrary scalar field dynamics. After changing to conformal time, $`\eta =𝑑t/a(t)`$, and introducing the canonical Mukhanov variable $`u^I=aQ^I`$, one obtains a matrix field equation: $$u^{I\prime \prime }+\left(\left[k^2\frac{a^{\prime \prime }}{a}\right]\delta _J^I+_{}^{I}{}_{J}{}^{}a^2\right)u^J=0,$$ (46) where a prime denotes derivatives with respect to $`\eta `$. Since the mass term $`_{}^{I}{}_{J}{}^{}`$ is not in general diagonal, Eq. (46) does not necessarily factorize into $`𝒩`$ independent equations for the $`u^I`$. Instead, the interactions described by $`_{}^{I}{}_{J}{}^{}`$ typically couple the scalar fields to one another. In the single-field scenario, the mass-matrix $`_{}^{I}{}_{J}{}^{}`$ can be expressed as a combination of slow-roll parameters. It can therefore be neglected to leading order in slow-roll. To arrive at a similar result in the multi-field case, one can express $`_{IJ}`$ directly in terms of the slow-roll matrices (10)–(16). It follows that $$\frac{1}{H^2}_{IJ}=3\stackrel{~}{\eta }_{IJ}6\epsilon _{IJ}\epsilon \epsilon _{IJ}+\epsilon \eta _{IJ}$$ (47) and Eq. (47) generalizes the familiar result from single-field inflation. Since we are only interested in leading order effects, we can consistently neglect the slow-roll contributions arising from $`_{IJ}`$ and consider only terms of $`\mathrm{O}(1)`$. This is the standard approximation which is invoked when estimating the amplitude and spectral index of inflationary fluctuations to lowest order in slow-roll , and when calculating the leading-order effects of non-gaussianity . It results in the de Sitter–Mukhanov equation: $$u^{I\prime \prime }+\left(k^2\frac{a^{\prime \prime }}{a}\right)u^I=u^{I\prime \prime }+\left(k^2\frac{2}{\eta ^2}\right)u^I=0.$$ (48) Applying the above simplification implies that we can work with the propagator $$Q^I(x_1)Q^J(x_2)=𝒢^{IJ}G_{}(x_1,x_2),$$ (49) where $`G_{}`$ represents the conventional scalar de Sitter Green’s function for a massless scalar field : $$G_{}(x_1,x_2)=\frac{H^2}{2k^3}\times \{\begin{array}{cc}(1\mathrm{i}k\eta _2)(1+\mathrm{i}k\eta _1)\mathrm{e}^{\mathrm{i}k(\eta _1\eta _2)}\hfill & \eta _1>\eta _2\hfill \\ (1\mathrm{i}k\eta _1)(1+\mathrm{i}k\eta _2)\mathrm{e}^{\mathrm{i}k(\eta _1\eta _2)}\hfill & \eta _2>\eta _1\hfill \end{array}$$ (50) By taking the coincidence limit of the propagator, we obtain the familiar power spectrum on large scales (where $`k0`$): $$Q^I(𝐤_1)Q^J(𝐤_2)=(2\pi )^3\delta (𝐤_1+𝐤_2)𝒢^{IJ}\frac{H^2}{2k^3},$$ (51) and it therefore follows that $$\mathrm{\Delta }_{}^2=\frac{H^2}{4\pi ^2}.$$ (52) ### 3.3 The third-order theory The leading-order slow-roll term in the third-order action is given by<sup>4</sup><sup>4</sup>4This expression should be compared, for example, with Eq. (3.6) of , to which it reduces in the single-field case. $`S_3={\displaystyle }\mathrm{d}t\mathrm{d}^3xa^3({\displaystyle \frac{1}{a^2}}𝒢_{IJ}\dot{Q}^I\psi Q^J{\displaystyle \frac{1}{4H}}𝒢_{MN}\dot{\phi }^MQ^N𝒢_{IJ}\dot{Q}^I\dot{Q}^J`$ $`{\displaystyle \frac{1}{a^4}}{\displaystyle \frac{1}{4H}}𝒢_{MN}\dot{\phi }^MQ^N𝒢_{IJ}Q^IQ^J),`$ (53) where $`\psi `$ was defined in (3.1) and is given to leading order by $$^2\psi =\frac{a^2}{2H}𝒢_{IJ}\dot{\phi }^I\dot{Q}^J.$$ (54) In exact analogy with the single-field calculation, it proves most convenient to integrate by parts in the term containing $`\psi `$. One then finds that $`{\displaystyle dt\mathrm{d}^3xa𝒢_{IJ}\dot{Q}^I\psi Q^J}=`$ $`{\displaystyle dt\mathrm{d}^3x\left(\frac{1}{2}aH𝒢_{IJ}\psi Q^IQ^J\frac{1}{2}a𝒢_{IJ}\dot{\psi }Q^IQ^J+a𝒢_{IJ}\dot{Q}^I\psi ^2Q^J\right)}.`$ (55) Eq. (54) may then be differentiated to determine $`\dot{\psi }`$. Keeping only leading-order terms, we find that $$\dot{\psi }=a^2𝒢_{IJ}\dot{\phi }^I^2\dot{Q}^J\frac{a^2}{2H}𝒢_{IJ}\dot{\phi }^I^2\ddot{Q}^J.$$ (56) It should be emphasized that (56) should not be employed directly in the action (3.3) since it involves second derivatives of the canonical field $`Q^I`$ and therefore changes the order of the field equations. Instead, by neglecting subleading terms in slow-roll, one may use the first-order equation of motion $$\frac{1}{a^3}\frac{\delta L}{\delta Q^I}|_1=3H\dot{Q}_I\ddot{Q}_I+\frac{1}{a^2}^2Q_I+\mathrm{O}(\epsilon )$$ (57) to eliminate the second-derivative term $`\ddot{Q}^I`$ in (56). It follows that $$\dot{\psi }=a^2\dot{\phi }^I^2\dot{Q}^I\frac{a^2}{2H}\dot{\phi }^I^2\left(\frac{1}{a^3}\frac{\delta L}{\delta Q^I}|_13H\dot{Q}_I+\frac{1}{a^2}^2Q_I\right)+\mathrm{O}(\epsilon ).$$ (58) Substituting (58) into (3.3) and performing a further integration by parts where necessary then results in the equivalent third-order action $`S_3={\displaystyle }\mathrm{d}t\mathrm{d}^3x({\displaystyle \frac{a^3}{4H}}\dot{\phi }^JQ_J\dot{Q}_I\dot{Q}^I{\displaystyle \frac{a^3}{2H}}\dot{\phi }^J^2\dot{Q}_J\dot{Q}_I^2Q^I`$ $`+{\displaystyle \frac{\delta L}{\delta Q_I}}|_1[{\displaystyle \frac{1}{4H}}\dot{\phi }^J𝒢_{IJ}^2(Q^K^2Q_K){\displaystyle \frac{1}{8H}}\dot{\phi }^J𝒢_{IJ}Q^KQ_K]).`$ (59) The last term in (59), which is proportional to the first-order equation of motion $`\delta L/\delta Q^I|_1`$, is familiar from the single-field calculation, and represents a field-redefinition. It can be eliminated from the action by transforming the fields $`Q^I`$ to new fields $`𝒬^I`$, which satisfy $$Q^I=𝒬^IF^I(Q)=𝒬^I\left(\frac{1}{4H}\dot{\phi }^I^2\left(𝒬^J^2𝒬_J\right)\frac{1}{8H}\dot{\phi }^I𝒬^J𝒬_J\right),$$ (60) where it makes no difference to leading order whether we write $`Q^J`$ or $`𝒬^J`$ in the quadratic terms. Since we are keeping only terms up to and including third order in $`Q`$, this redefinition has no effect on the interactions described in (59), and we may freely set $`Q^I=𝒬^I`$ there. However, the redefinition does modify the quadratic part of the action, Eq. (44), which transforms as $$S_2[Q]S_2[𝒬]dt\mathrm{d}^3x𝒢_{IJ}\frac{\delta L}{\delta Q^I}|_1F^J.$$ (61) The extra term here exactly suffices to cancel the interactions in (59) which were proportional to the first-order equations of motion, $`\delta L/\delta Q^I|_1`$. Although this field redefinition is extremely convenient for the purposes of calculation, we are ultimately interested in the correlators of $`Q^I`$ and not $`𝒬^I`$. The correlators are related via the standard prescription $`Q^IQ^JQ^K={\displaystyle [\mathrm{d}Q^M]Q^IQ^JQ^J\mathrm{e}^{\mathrm{i}S[Q]}}`$ $`{\displaystyle [\mathrm{d}𝒬^M]\left(𝒬^I𝒬^J𝒬^K+F^I𝒬^J𝒬^K+\text{cyclic}\right)\mathrm{e}^{\mathrm{i}S[𝒬]}}+\mathrm{O}(\epsilon ),`$ (62) which is just Wick’s theorem. This allows us to add in the contribution of the redefined terms in (60) at the end of the calculation. In the present calculation, the very complicated field redefinition found by Maldacena (displayed in Eq. (3.8) of Ref. ) does not appear. This is because such a redefinition of fields is almost exactly equivalent to a translation from the comoving to the uniform-curvature gauge. In this paper, we have worked in the uniform-curvature gauge from the outset, which results in a considerable simplification even for the example of a single field. In some sense, this effect demonstrates that the uniform-curvature gauge is especially convenient for calculations of this sort: in the single-field case, one is free to work equally in the comoving gauge (as in Refs. ) or the uniform-curvature gauge, but in order to actually perform the functional integrals implicit in the three-point correlator $`\zeta \zeta \zeta `$, one must eventually return to the uniform-curvature gauge, except for an anomalous field redefinition related to the effect of the $`\ddot{Q}^I`$ term. It is interesting to note that if we had naïvely kept $`\ddot{Q}^I`$ in the action, instead of replacing it with the first-order equation of motion, we would have omitted the contribution of the quadratic terms in (60). This may seem paradoxical, but, as we pointed out above, the appearance of $`\ddot{Q}^I`$ in the action changes the order of the field equations, and consequently their dynamical behaviour. We wish to retain only perturbations around the branch of solutions described by (3), and no others. For this reason, we are obliged to eliminate any second- or higher-derivative terms which appear in the slow-roll expansion of the action for the $`Q^I`$, and we achieve this by employing the background attractor behaviour of the second-order action to select the correct solution around which we perturb. ### 3.4 Calculating the three-point function We are now in a position to calculate the three-point function $`𝒬^I(𝐤_1)𝒬^J(𝐤_2)𝒬^K(𝐤_3)`$ corresponding to the action (59), where we work with the redefined field $`𝒬^I`$, defined in Eq. (60), so that the $`\delta L/\delta Q_I|_1`$ terms are not present. A detailed description of how the three-point calculation is carried out has been presented elsewhere , so here we merely record the results. * The first interaction to consider is given by $$d\eta \mathrm{d}^3x\left(\frac{a^2}{4H}\dot{\phi }^J𝒬_J𝒬_I^{}𝒬^I\right),$$ (63) where we have converted the integral to conformal time. This gives a contribution to $`𝒬^I𝒬^J𝒬^K`$, after translation to Fourier space, of the form $$\mathrm{i}(2\pi )^3\delta (\underset{i}{}𝐤_i)\frac{H^3}{4}\frac{1}{_i2k_i^3}\dot{\phi }^I𝒢^{JK}d\eta k_2^2k_3^2(1\mathrm{i}k_1\eta )\mathrm{e}^{\mathrm{i}k_t\eta }+\text{perms},$$ (64) where $`k_t=k_1+k_2+k_3`$ is the total (scalar) momentum. The integral can be evaluated after Wick rotation onto the positive imaginary axis and it then follows that this contribution is equivalent to $$(2\pi )^3\delta (\underset{i}{}𝐤_i)\frac{H_{}^3}{4}\frac{1}{_i2k_i^3}\left[\underset{\text{perms}}{}\dot{\phi }_{}^I𝒢^{JK}\left(\frac{k_2^2k_3^2}{k_t}\frac{k_1k_2^2k_3^2}{k_t^2}\right)\right],$$ (65) where an asterisk ‘$``$’ denotes evaluation at horizon crossing. The sum in this and subsequent expressions is over all ways of simultaneously rearranging the indices $`I`$, $`J`$ and $`K`$ and the momenta $`k_1`$, $`k_2`$ and $`k_3`$, such that the relative positioning of the $`k`$’s is respected. (In other words, when exchanging indices $`I`$ and $`J`$, for example, one should also exchange $`k_1`$ and $`k_2`$, and so on.) * The second relevant interaction is $$d\eta \mathrm{d}^3x\left(\frac{a^2}{2H}\dot{\phi }^J^2𝒬_J^{}𝒬_I^{}^2𝒬^I\right),$$ (66) where again we have rewritten the integral in terms of conformal time. After translating to Fourier space and Wick rotating the resulting integral, we have that $$(2\pi )^3\delta (\underset{i}{}𝐤_i)\frac{H_{}^3}{2}\frac{1}{_i2k_i^3}\left[\underset{\text{perms}}{}\dot{\phi }_{}^I𝒢^{JK}\left(\frac{k_2^2k_3^2}{k_t}\frac{k_2^2k_3^3}{k_t^2}\right)\right].$$ (67) Once the field redefinition terms in (59) have been introduced back into the three-point function, using (62), one finds that $$Q^I(𝐤_1)Q^J(𝐤_2)Q^K(𝐤_3)=(2\pi )^3\delta (\underset{i}{}𝐤_i)\frac{4\pi ^4}{_ik_i^3}|\mathrm{\Delta }_{}^2|^2𝒜^{IJK}(k_1,k_2,k_3),$$ (68) where the spectrum $`\mathrm{\Delta }_{}^2`$ is to be evaluated at horizon crossing, and $`𝒜^{IJK}`$ is a momentum-dependent function given by $`𝒜^{IJK}(k_1,k_2,k_3)={\displaystyle \underset{\text{perms}}{}}{\displaystyle \frac{\dot{\phi }_{}^I}{4H_{}}}𝒢^{JK}\left(3{\displaystyle \frac{k_2^2k_3^2}{k_t}}{\displaystyle \frac{k_2^2k_3^2}{k_t^2}}(k_1+2k_3)+{\displaystyle \frac{1}{2}}k_1^3k_1k_2^2\right)`$ $`={\displaystyle \underset{\text{perms}}{}}{\displaystyle \frac{1}{2\sqrt{2}}}\epsilon _{}^I𝒢^{JK}\left(3{\displaystyle \frac{k_2^2k_3^2}{k_t}}{\displaystyle \frac{k_2^2k_3^2}{k_t^2}}(k_1+2k_3)+{\displaystyle \frac{1}{2}}k_1^3k_1k_2^2\right)`$ (69) This is of order a slow-roll parameter, so (as we expect) the non-gaussianity will not be large when $`|\epsilon ^{IJ}|1`$. Eqs. (68)–(69) represent the main result of this paper. The Feynman integrals that lead to $`Q^IQ^JQ^K`$ are insensitive to the field modes’ behaviour deep inside the horizon and long after the modes have passed outside the horizon. The dominant contribution arises from fluctuations around the epoch of horizon crossing . For this reason, (68) represents the primordial three-point correlation among the scalars $`Q^I`$ a short time after horizon crossing. In particular, in making the estimate (68)–(69), we have assumed that the three $`𝐤`$-modes have roughly comparable wavenumbers, so that they cross the horizon at similar epochs. Eqs. (68)–(69) do not by themselves represent the primordial non-gaussianity; instead, they only describe cubic interactions among the $`Q^I`$ at horizon crossing. In order to use Eqs. (68)–(69) to calculate the primordial non-gaussianity in the curvature perturbation, it is necessary to use the general formalism outlined in Section 2 to assemble these initial non-gaussianities into the actual non-gaussianities that are in principle observable in the CMB. In the following sections, we consider two specific scenarios in which we can carry out this detailed assembly explicitly. We first verify that our results reduce to the well-known result of single-field inflation. We then proceed in Section 5 investigate the multiple-field analogue of the single-field power-law model known as assisted inflation . ## 4 Reduction to single-field case In this section we verify that Eq. (68) reduces to the result of Maldacena for the case of inflation driven by a single scalar field. We work with the large-scale expression (23), which requires an explicit expression for the number of e-folds, $`N`$, as a function of the field. For a single field $`\phi `$ with perturbation $`Q`$ the integrated e-folding rate is given by (17): $$N=Hdt=\frac{1}{\epsilon }\mathrm{ln}H$$ (70) whenever $`\epsilon `$ is approximately constant over the range of e-folds under consideration. Therefore, the derivatives of $`N`$, which are needed to determine $`\zeta `$, are $$N_{,\phi }=\frac{1}{\epsilon }\frac{H_{,\phi }}{H},N_{,\phi \phi }=\frac{1}{\epsilon }\frac{H_{,\phi \phi }}{H}+\frac{1}{\epsilon }\left(\frac{H_{,\phi }}{H}\right)^2.$$ (71) Recalling that for single field models, the slow-roll parameters can be expressed as $$\epsilon =2\left(\frac{H_{,\phi }}{H}\right)^2,\eta =\frac{\ddot{\phi }}{H\dot{\phi }}=2\frac{H_{,\phi \phi }}{H},$$ (72) we may write $`\zeta `$ as $$\zeta =\frac{1}{\sqrt{2\epsilon }}Q+\left(\frac{1}{4}\frac{\eta }{4\epsilon }\right)Q^2.$$ (73) As expected, to lowest-order this implies that the spectra are related by the well-known result $$\zeta (𝐤_1)\zeta (𝐤_2)=\frac{1}{2\epsilon }Q(𝐤_1)Q(𝐤_2),$$ (74) with $`\epsilon `$ evaluated at the moment of horizon crossing. A similar expression holds for the three-point function. It follows from (62) that the $`\zeta \zeta \zeta `$ and $`QQQ`$ correlators are related via $`\zeta (𝐤_1)\zeta (𝐤_2)\zeta (𝐤_3)={\displaystyle \frac{1}{(2\epsilon )^{3/2}}}Q(𝐤_1)Q(𝐤_2)Q(𝐤_3)`$ $`+{\displaystyle \frac{1}{2\epsilon }}\left({\displaystyle \frac{1}{4}}{\displaystyle \frac{\eta }{4\epsilon }}\right)Q(𝐤_1)Q(𝐤_2)[QQ](𝐤_3)+\text{perms},`$ (75) where (as before) $``$ denotes a convolution product. Eq. (4) implies that the *connected* part of the $`\zeta `$ three-point function is related to the connected $`Q`$ three-point function by $$\zeta (𝐤_1)\zeta (𝐤_2)\zeta (𝐤_3)_c=\frac{H_{}^3}{\dot{\phi }_{}^3}Q(𝐤_1)Q(𝐤_2)Q(𝐤_3)_c$$ (76) where a subscript $`c`$ denotes the connected part of a correlation function. However, the prescription of (75) means that we must mix some disconnected pieces with this connected component, which are described by the four-term correlators $`QQQQ`$. These are no more than the superhorizon parts of the gauge transformation between the comoving and uniform curvature gauges, which were derived by Maldacena and found to have the form $$(2\pi )^3\delta (\underset{i}{}𝐤_i)\frac{H^4}{4}\left(\frac{1}{2}\frac{H^4}{\dot{\phi }^4}\frac{\ddot{\phi }}{H\dot{\phi }}+\frac{1}{4}\frac{H^2}{\dot{\phi }^2}\right)\frac{1}{k_2^2k_3^3}+\text{perms}.$$ (77) Thus, the complete three-point function which we have calculated via this method is $$\zeta (𝐤_1)\zeta (𝐤_2)\zeta (𝐤_3)=(2\pi )^3\delta (\underset{i}{}𝐤_i)\frac{H_{}^8}{\dot{\phi }_{}^4}\frac{1}{_i2k_i^3}𝒜,$$ (78) where $`𝒜`$ is given by $$𝒜=\frac{\dot{\phi }_{}^2}{H_{}^2}\frac{4}{k_t}\underset{i>j}{}k_i^2k_j^2+\left(2\frac{\ddot{\phi }_{}}{H_{}\dot{\phi }_{}}+\frac{1}{2}\frac{\dot{\phi }_{}^2}{H_{}^2}\right)\underset{i}{}k_i^3+\frac{1}{2}\frac{\dot{\phi }_{}^2}{H_{}^2}\underset{ij}{}k_ik_j^2,$$ (79) in exact agreement with Maldacena’s result. ## 5 Assisted inflation In this section we calculate the three-point correlation function for a multiple-field model of inflation known as assisted inflation . In this scenario, a set of $`𝒩`$ scalar fields, each with an exponential potential, can conspire to drive a phase of accelerated expansion, even if the individual potential for each field is too steep to support inflation. The potentials are given by $$V_I(\phi ^I)=V_0\mathrm{exp}\left(\sqrt{\frac{2}{p^I}}\phi ^I\right),$$ (80) where $`V_0`$ is a constant that determines the overall scale of the model, and $`p^I`$ represent coupling constants.<sup>5</sup><sup>5</sup>5The summation convention for indices $`I`$, …, $`J`$ is suspended in this section. In the single field model, $`p^1>1`$ is required for inflation. The combined effective potential is given by $`V=_IV_I`$. In the spatially flat FRW cosmology, the system of fields (80) admit a late-time attractor solution characterized by $$\phi ^I=\sqrt{\frac{p^I}{p^1}}\phi ^1+\alpha ^I,$$ (81) for some constants $`\alpha ^I`$ and any choice of reference field, which we have arbitrarily identified as $`\phi ^1`$. At the classical level, the existence of this attractor solution implies that the model is dynamically equivalent to that of a single scalar field with an exponential potential characterized by a coupling constant $`\stackrel{~}{p}=_Ip^I`$. To proceed, we need to determine the number of e-folds in terms of the scalar field values. Following a standard calculation , we can express $`\mathrm{d}N`$ as $$\mathrm{d}N=H\frac{\mathrm{d}\phi _I}{\dot{\phi }_I},$$ (82) for any field $`\phi _I`$. When slow-roll is valid, we can also make use of the approximate field equations $`3H\dot{\phi }_I+\mathrm{d}V_I/\mathrm{d}\phi _I=0`$, such that $$\mathrm{d}N\frac{3H^2}{\mathrm{d}V_1/\mathrm{d}\phi _1}\mathrm{d}\phi _1\frac{V}{\mathrm{d}V_1/\mathrm{d}\phi _1}\mathrm{d}\phi _1=\underset{I}{}\frac{V_I}{\mathrm{d}V_I/\mathrm{d}\phi _I}\mathrm{d}\phi _I.$$ (83) The first and second derivatives of $`N`$ now follow immediately from Eq. (83): $`{\displaystyle \frac{\mathrm{d}N}{\mathrm{d}\phi _I}}={\displaystyle \frac{V_I}{\mathrm{d}V_I/\mathrm{d}\phi _I}}`$ $`{\displaystyle \frac{\mathrm{d}^2N}{\mathrm{d}\phi _I^2}}=1V_I{\displaystyle \frac{\mathrm{d}^2V_I}{\mathrm{d}\phi _I^2}}\left({\displaystyle \frac{\mathrm{d}V_I}{\mathrm{d}\phi _I}}\right)^2.`$ (84) Moreover, for exponential potentials it follows that $$\frac{V_I}{\mathrm{d}V_I/\mathrm{d}\phi _I}=\sqrt{\frac{p^I}{2}},\frac{\mathrm{d}^2V_I/\mathrm{d}\phi _I^2}{\mathrm{d}V_I/\mathrm{d}\phi _I}=\sqrt{\frac{2}{p^I}}.$$ (85) Hence, the second derivative of $`N`$ vanishes and the only remaining contribution to the total curvature perturbation (23) is provided by the linear derivative term, which simplifies to $$\zeta =\underset{I}{}\sqrt{\frac{p^I}{2}}Q^I.$$ (86) At the level of the two-point correlation function, this implies that $$\zeta (𝐤_1)\zeta (𝐤_2)=\underset{I,J}{}\frac{\sqrt{p^Ip^J}}{2}Q^IQ^J=(2\pi )^3\delta (𝐤_1+𝐤_2)\stackrel{~}{p}\frac{H_{}^2}{4k_1^3}$$ (87) (with an asterisk ‘$``$’ indicating evaluation at horizon exit, as usual) and this is precisely the spectrum that would arise for a single-field model with an exponential potential characterized by $`\stackrel{~}{p}`$, in agreement with the classical observation that assisted inflation is dynamically equivalent to such a model . On the other hand, we find that the three-point function is given by $$\zeta (𝐤_1)\zeta (𝐤_2)\zeta (𝐤_2)=\underset{I,J,L}{}\sqrt{\frac{p^Ip^Jp^K}{8}}(2\pi )^3\delta (\underset{i}{}𝐤_i)\frac{H_{}^4}{4}\frac{1}{_ik_i^3}𝒜^{IJK}(k_1,k_2,k_3).$$ (88) In order to correctly estimate this amplitude, we need an expression for the slow-roll parameters. We take $$\epsilon ^I=\frac{\sqrt{p^I}}{\stackrel{~}{p}}.$$ (89) Thus, $`\mathrm{tr}\epsilon ^{IJ}=1/\stackrel{~}{p}`$, which coincides with the slow-roll parameter in the classically equivalent single-field model. After some algebra it can be shown that this is equivalent to $$\zeta (𝐤_1)\zeta (𝐤_2)\zeta (𝐤_2)=(2\pi )^3\delta (\underset{i}{}𝐤_i)\frac{H_{}^4}{_i2k_i^3}\frac{1}{2}\stackrel{~}{p}\left(\frac{4}{k_t}\underset{i>j}{}k_i^2k_j^2\frac{1}{2}\underset{i}{}k_i^3+\frac{1}{2}\underset{ij}{}k_ik_j^2\right),$$ (90) which is identical to the equivalent single-scalar model evaluated with Maldacena’s single-field formula. This reduction to the single-field result illustrates a general feature of multiple-field models. In what amounts to the familiar Gram–Schmidt procedure, one can always make a rotation in field space so that one distinguished field (say $`Q^1`$) lies along the adiabatic direction, and the remaining $`𝒩1`$ fields describe isocurvature fluctuations around it . These isocurvature fluctuations do not contribute to the spectrum or bispectrum. Therefore, quite generally, one expects the resultant three-point function to be given by Maldacena’s single-field result, constructed using the field-space trajectory of the adiabatic field. One should not understand this to mean that the formalism outlined in this paper is unnecessary. On the contrary, although the rotation in field space which produces an adiabatic field and $`𝒩1`$ isocurvature fields can be performed explicitly in simple cases , for a completely general multiple-field model, the most convenient way to account for the effect of this rotation (at least on large scales) is to use exactly the non-linear extension of the Sasaki–Stewart $`\delta N`$ formalism, which was described in Section 2.2. In this formalism it will always be necessary to ‘prime’ the superhorizon evolution using what amounts to an initial condition, describing the non-gaussianity produced at horizon crossing. Eq. (69) constitutes exactly the necessary initial condition. ## 6 Conclusions In this paper, we have calculated the three-point function evaluated just after horizon exit that arises from a set of $`𝒩`$ scalar field fluctuations $`Q^I`$ coupled to gravity during inflation. The result is given by Eq. (38), where the momentum-dependent contribution $`𝒜^{IJK}`$ is given in Eqs. (68)–(69). This three-point function measures the intrinsic non-gaussianity produced at horizon exit in the coupled scalar–gravity system, and is important since it provides the initial condition needed for a calculation of the superhorizon evolution of the non-gaussianity in the inflationary density perturbation . Such superhorizon physics describes how the curvature perturbation $`\zeta `$, which is used to calculate observable microwave background anisotropies via the Sachs–Wolfe effect, evolves outside the Hubble radius. Even though the initial non-gaussianity we have calculated, which is set at horizon exit, is typically small, the total non-gaussianity of $`\zeta `$ after superhorizon evolution can be large. For example, this can occur in curvaton-like scenarios or where there is cross-correlation between the adiabatic and isocurvature modes . As a result, in some circumstances our expression could be used to verify that the non-gaussian effects generated at horizon crossing by microphysical processes are subdominant to the superhorizon, isocurvature-driven evolution of $`\zeta `$, as described in . However, more generally, both of these effects will be important and our result contributes in an essential way to the primordial non-gaussianity produced by inflation. It is important to note that (68)–(69) describe the non-linearities among the $`Q^I`$ at horizon crossing. The details of this were recently explored by Weinberg , who showed that in a large class of multi-field models such integrals are dominated by the epoch of horizon-crossing, even beyond tree-level in perturbation theory. Provided one evaluates the slow-roll parameters when the $`𝐤`$-modes in question crossed the horizon, Eq. (69) gives the relevant non-linearities. Of course, both the $`Q^I`$ and $`\zeta `$ continue to evolve on superhorizon scales. For this reason one must assemble the $`QQQ`$ correlators into $`\zeta \zeta \zeta `$ correlators to follow the time dependence correctly when outside the Hubble radius. The same conclusion was reached by Rigopoulos, Shellard & van Tent . We do not address the question of what happens when one $`𝐤`$-mode is squeezed, so that $`k_1k_2,k_3`$ (say). In this case the mode corresponding to $`𝐤_1`$ is pushed outside the horizon much earlier than the other two, since it is comparatively larger. In the single-field case, where perturbations “freeze in” outside the horizon owing to the constancy of $`\zeta `$, the three-point function factorizes into the two-point function of $`𝐤_2`$ and $`𝐤_3`$ evaluated in a background which takes into account the back-reaction of the frozen $`𝐤_1`$ perturbation . In a multiple-field model, the $`𝐤_1`$ mode will come to constitute part of the zero-momentum background when it is far outside the horizon, but since $`\zeta `$ is no longer constant it is not so straightforward to provide a quantitative description of its effect. We leave this interesting question for future work. The level of non-gaussianity arising from microphysics is typically proportional to a slow-roll parameter, $`\epsilon ^{1/2}𝒜^{IJK}`$, and is therefore expected to be small when fluctuations are close to scale-invariance. Consequently, when isocurvature modes drive a very strong superhorizon evolution of $`\zeta `$, and thereby source considerable non-gaussianity on very large scales, it seems likely that the generic situation will match that described by Lyth & Rodriguez , although the only model we are aware of where this can be verified in detail is the curvaton scenario. On the other hand, if the superhorizon contribution never becomes large, the non-gaussian effects we have considered will be important, in which case they can not be neglected. This is the situation that arises, for example, in the assisted inflationary scenario considered in Section 5, which is the canonical model where the superhorizon piece is entirely absent. It is worth emphasizing that the three-point function can be calculated once the evolution of the unperturbed background cosmology, as parametrized by $`N(\phi ^I)`$, has been determined. Our result therefore complements that of , and implies that whenever one has enough information about the background dynamics to calculate the momentum-independent piece of $`f_{\mathrm{NL}}`$, there will be always be enough information to calculate the momentum-dependent contribution as well. In this sense, no further work is required in order to calculate the full, momentum-dependent $`f_{\mathrm{NL}}`$. In general, the structure of the three-point function (68)–(69) is sensitive to the target space metric $`𝒢^{IJ}`$. A similar dependence arises in the two-point function, which is proportional to $`𝒢^{IJ}`$. Indeed, neglecting numerical coefficients, the form of (69) is in some sense inevitable. Specifically, it must arise from a cubic combination of the $`Q^A`$’s which is a singlet with respect to the target space metric. Since the $`Q^A`$ can only appear contracted with the metric $`𝒢_{IJ}`$, it follows that to leading-order in slow-roll, the only possible combination must have the form $`𝒢_{AB}\dot{\phi }^AQ^B𝒢_{IJ}Q^IQ^J`$. Such a combination leads to the index structure appearing in (69), there being no other target space vector at lowest-order in slow-roll which can be contracted with $`G_{AB}Q^B`$ to yield a singlet. It is natural to speculate that an analogous result will hold for the general $`n`$-th–order correlation function. When $`n`$ is even ($`n=2m`$) we expect the leading order slow-roll contribution to arise from a scalar quantity containing $`m`$ contractions with $`𝒢_{IJ}`$: $$\underset{2m\text{ copies}}{\underset{}{Q^AQ^B\mathrm{}Q^EQ^F}}\left(\frac{\dot{\phi }}{H}\right)^{2(m1)}\underset{m\text{ copies}}{\underset{}{𝒢^{(AB}\mathrm{}𝒢^{EF)}}},$$ (91) where we have neglected momentum-dependent factors which may accompany each factor of the target space metric $`𝒢^{AB}`$. When $`n`$ is odd, so that $`n=2m+1`$, we anticipate instead that $$\underset{2m+1\text{ copies}}{\underset{}{Q^AQ^BQ^C\mathrm{}Q^EQ^F}}\left(\frac{\dot{\phi }}{H}\right)^{2(m1)}H^1\dot{\phi }^{(A}\underset{m\text{ copies}}{\underset{}{𝒢^{BC}\mathrm{}𝒢^{EF)}}}.$$ (92) We have restricted our attention to the case where the target-space metric is flat and independent of the scalar field values, $`𝒢^{IJ}=\delta ^{IJ}`$. Although this Ansatz covers a very wide class of multi-field inflationary models, it would be interesting to go beyond this approximation (see, for example, ). Such an extension has been considered recently using a stochastic approach , although in principle, more general scenarios of this type could be investigated by employing the formalism developed in the present paper. For example, a specific point of interest concerns the curvature of the scalar field manifold. In general, a non-trivial metric $`𝒢_{IJ}(\phi )`$ will result in a curvature tensor of the form $$\mathrm{\Omega }_{}^{A}{}_{BCD}{}^{}=\omega _{}^{A}{}_{BD,C}{}^{}\omega _{}^{A}{}_{BC,D}{}^{}+\omega _{}^{F}{}_{BD}{}^{}\omega _{}^{A}{}_{FC}{}^{}\omega _{}^{F}{}_{BC}{}^{}\omega _{}^{A}{}_{FD}{}^{},$$ (93) where $`\omega _{}^{A}{}_{BC}{}^{}`$ is the Levi-Civita connexion compatible with $`𝒢_{IJ}`$. This tensor is identically zero for the models considered in this paper, but is expected to arise in the expressions for the spectral tilt and other cosmological observables in more general classes of models. It is possible that the presence of curvature terms of this form may be sufficiently important to invalidate the flat target-space analysis we have adopted. In this case, a larger primordial non-gaussianity could be generated. However, since the curvature tensor $`\mathrm{\Omega }_{}^{A}{}_{BCD}{}^{}`$ can not be too large if slow-roll is to be respected , we anticipate that this will not be the case in general. DS is supported by PPARC. We would like to thank David Lyth and David Wands for useful discussions. ## References